arxiv: v2 [astro-ph.ga] 1 Apr 2018

Size: px
Start display at page:

Download "arxiv: v2 [astro-ph.ga] 1 Apr 2018"

Transcription

1 Extreme magnification of a star at redshift 1.5 by a galaxycluster lens Patrick L. Kelly 1,2, Jose M. Diego 3, Steven Rodney 4, Nick Kaiser 5, Tom Broadhurst 6,7, Adi Zitrin 8, Tommaso Treu 9, Pablo G. Pérez-González 10, Takahiro Morishita 9,11,12, Mathilde Jauzac 13,14,15, arxiv: v2 [astro-ph.ga] 1 Apr 2018 Jonatan Selsing 16, Masamune Oguri 17,18,19, Laurent Pueyo 20, Timothy W. Ross 1, Alexei V. Filippenko 1,21, Nathan Smith 22, Jens Hjorth 16, S. Bradley Cenko 23,24, Xin Wang 9, D. Andrew Howell 25,26, Johan Richard 27, Brenda L. Frye 22, Saurabh W. Jha 28, Ryan J. Foley 29, Colin Norman 30, Marusa Bradac 31, Weikang Zheng 1, Gabriel Brammer 20, Alberto Molino Benito 32, Antonio Cava 33, Lise Christensen 16, Selma E. de Mink 34, Or Graur 35,36,37, Claudio Grillo 38,16, Ryota Kawamata 39, Jean- Paul Kneib 40, Thomas Matheson 41, Curtis McCully 25,26, Mario Nonino 42, Ismael Perez-Fournon 43,44, Adam G. Riess 30,20, Piero Rosati 45, Kasper Borello Schmidt 46, Keren Sharon 47, & Benjamin J. Weiner 22 1 Department of Astronomy, University of California, Berkeley, CA , USA 2 School of Physics and Astronomy, University of Minnesota, 116 Church Street SE, Minneapolis, MN 55455, USA 3 IFCA, Instituto de Física de Cantabria (UC-CSIC), Av. de Los Castros s/n, Santander, Spain 4 Department of Physics and Astronomy, University of South Carolina, 712 Main St., Columbia, SC 29208, USA 5 Institute for Astronomy, University of Hawaii, 2680 Woodlawn Drive, Honolulu, HI , USA 1

2 6 Department of Theoretical Physics, University of the Basque Country, Bilbao 48080, Spain 7 IKERBASQUE, Basque Foundation for Science, Alameda Urquijo, Bilbao, Spain 8 Physics Department, Ben-Gurion University of the Negev, P.O. Box 653, beer-sheva , Israel 9 Department of Physics and Astronomy, University of California, Los Angeles, CA Departamento de Astrofísica, Facultad de critical curve. Físicas, Universidad Complutense de Madrid, E Madrid, Spain 11 Astronomical Institute, Tohoku University, Aramaki, Aoba, Sendai , Japan 12 Institute for International Advanced Research and Education, Tohoku University, Aramaki, Aoba, Sendai , Japan 13 Centre for Extragalactic Astronomy, Department of Physics, Durham University, Durham DH1 3LE, U.K. 14 Institute for Computational Cosmology, Durham University, South Road, Durham DH1 3LE, U.K. 15 Astrophysics and Cosmology Research Unit, School of Mathematical Sciences, University of KwaZulu-Natal, Durban 4041, South Africa 16 Dark Cosmology Centre, Niels Bohr Institute, University of Copenhagen, Juliane Maries Vej 30, DK-2100 Copenhagen, Denmark 17 Research Center for the Early Universe, University of Tokyo, Tokyo , Japan 18 Department of Physics, University of Tokyo, Hongo, Bunkyo-ku, Tokyo , Japan 19 Kavli Institute for the Physics and Mathematics of the Universe (Kavli IPMU, WPI), University 2

3 of Tokyo, Kashiwanoha, Kashiwa, Chiba , Japan 20 Space Telescope Science Institute, 3700 San Martin Dr., Baltimore, MD 21218, USA 21 Miller Senior Fellow, Miller Institute for Basic Research in Science, University of California, Berkeley, CA 94720, USA 22 Steward Observatory, University of Arizona, 933 N. Cherry Ave., Tucson, AZ 85721, USA 23 Astrophysics Science Division, NASA Goddard Space Flight Center, Mail Code 661, Greenbelt, MD 20771, USA 24 Joint Space-Science Institute, University of Maryland, College Park, MD 20742, USA 25 Las Cumbres Observatory, 6740 Cortona Dr., Suite 102, Goleta, CA 93117, USA 26 Department of Physics, University of California, Santa Barbara, CA , USA 27 Univ Lyon, Univ Lyon1, ENS de Lyon, CNRS, Centre de Recherche Astrophysique de Lyon UMR5574, F-69230, Saint-Genis-Laval, France 28 Department of Physics and Astronomy, Rutgers, The State University of New Jersey, Piscataway, NJ 08854, USA 29 Department of Astronomy and Astrophysics, UCO/Lick Observatory, University of California, 1156 High Street, Santa Cruz, CA 95064, USA 30 Department of Physics and Astronomy, The Johns Hopkins University, 3400 N. Charles St., Baltimore, MD 21218, USA 31 Department of Physics, University of California, Davis, 1 Shields Avenue, Davis, CA 95616, USA 32 Instituto de Astronomia, Geofísica e Ciências Atmosféricas, Universidade de São Paulo,

4 090, São Paulo, Brazil 33 Department of Astronomy, University of Geneva, 51, Ch. des Maillettes, CH-1290 Versoix, Switzerland 34 Anton Pannekoek Institute for Astronomy, University of Amsterdam, NL-1090 GE Amsterdam, the Netherlands 35 Harvard-Smithsonian Center for Astrophysics, 60 Garden Street, Cambridge, MA Department of Astrophysics, American Museum of Natural History, Central Park West and 79th Street, New York, NY 10024, USA 37 NSF Astronomy and Astrophysics Postdoctoral Fellow 38 Dipartimento di Fisica, Universitá degli Studi di Milano, via Celoria 16, I Milano, Italy 39 Department of Astronomy, Graduate School of Science, The University of Tokyo, Hongo, Bunkyo-ku, Tokyo , Japan 40 Laboratoire d Astrophysique, Ecole Polytechnique Federale de Lausanne (EPFL), Observatoire de Sauverny, CH-1290 Versoix, Switzerland 41 National Optical Astronomical Observatory, Tucson, AZ 85719, USA 42 INAF, Osservatorio Astronomico di Trieste, via Bazzoni 2, Trieste, Italy 43 Instituto de Astrofisica de Canarias (IAC), E La Laguna, Tenerife, Spain 44 Universidad de La Laguna, Dpto. Astrofisica, E La Laguna, Tenerife, Spain 45 Dipartimento di Fisica e Scienze della Terra, Universitá degli Studi di Ferrara, via Saragat 1, I-44122, Ferrara, Italy 46 Leibniz-Institut fur Astrophysik Potsdam (AIP), An der Sternwarte 16, Potsdam, Germany 4

5 47 University of Michigan, Department of Astronomy, 1085 South University Avenue, Ann Arbor, MI , USA Galaxy-cluster gravitational lenses can magnify background galaxies by a total factor of up to 50. Here we report an image of an individual star at redshift z = 1.49 (dubbed MACS J1149 Lensed Star 1 (LS1) ) magnified by > A separate image, detected briefly 0.26 from LS1, is likely a counterimage of the first star demagnified for multiple years by a 3 M object in the cluster. For reasonable assumptions about the lensing system, microlensing fluctuations in the stars light curves can yield evidence about the mass function of intracluster stars and compact objects, including binary fractions and specific stellar evolution and supernova models. Dark-matter subhalos or massive compact objects may help to account for the two images long-term brightness ratio. The pattern of magnification arising from a foreground strong gravitational lens changes with distance behind it. At each specific distance behind the lens, the locations that are most highly magnified are connected by a so-called caustic curve. Near the caustic curve in the source plane, magnification changes rapidly. Over a distance of only tens of parsecs close to the MACS J1149 galaxy cluster s caustic at z = 1.5, for example, magnification falls from a maximum of 5000 to only 50. Since the sizes of even compact galaxies are hundreds of parsecs, their total magnifications cannot exceed 50. However, a well-aligned individual star adjacent to the caustic of a galaxy cluster could, in theory, become magnified by a factor of many thousands 1. When a galaxy cluster s caustic curve 5

6 is mapped from the source plane defined at a specific redshift to the image plane on the sky, it is called the critical curve. Consequently, a highly magnified star should be found close to the foreground galaxy cluster s critical curve. 1 A Lensed Blue Supergiant at Redshift z = 1.49 In Hubble Space Telescope (HST) Wide Field Camera 3 (WFC3) infrared (IR) imaging taken on 29 April 2016 to construct light curves of the multiple images of Supernova (SN) Refsdal 2 11, we detected an unexpected change in flux of an individual point source (dubbed MACS J1149 Lensed Star 1 (LS1) ) in the MACS J1149 galaxy-cluster field 12. As shown in Fig. 1, the unresolved blue source lies close to the cluster s critical curve at its host galaxy s redshift of z = Fig. 2 shows that, while the location of the critical curve differs by 0.25 among lens models, the blue source is no farther than 0.13 from the critical curves of all publicly available, high-resolution models. The MACS J1149 galaxy cluster lens creates two partial, merging images of LS1 s host galaxy separated by the cluster critical curve, as well as an additional full image. As shown in Extended Data Fig. 3, LS1 s predicted position inside the third, full image is near the tip of a spiral arm. According to our lens model, LS1 is 7.9 ± 0.5 kpc from the nucleus of the host galaxy. The multiply imaged SN Refsdal exploded at a different position in the same galaxy At the peak of the microlensing event in May 2016 (Lensing Event Lev16A ), LS1 was a factor of 4 times brighter than it appeared in archival HST imaging during Fig. 3 6

7 shows that the additional flux we measured at LS1 s position has a spectral energy distribution (SED) statistically consistent with the source s SED during As shown in Fig. 3, model spectra of mid-to-late B-type stars at z = 1.49 with photospheric temperatures of 11,000 14,000 K 17 provide a good match to the SED of LS1 (χ 2 = 12.9 for 6 degrees of freedom; χ 2 ν = 2.15), given that our model does not account for changes in magnification between the epochs when observations in separate filters were obtained. SED fitting finds probability peaks at 8 and 35 Myr (see Extended Data Fig. 1) for the age of the arc underlying LS1 s position. A lensed luminous star provides a perhaps unexpected explanation (and yet the only reasonable one we could find) for the transient s variable light curve and unchanging SED. Except for finite-source effects, gravitational lensing will magnify a star s emission at all wavelengths equally. Therefore, as we observe for LS1, the SED of a lensed background star should remain the same, even as it appears brighter or fainter owing to changes in its magnification. By contrast, the SEDs of stellar outbursts and supernovae change as they brighten by the factor of 4 observed in May As shown in Fig. 3, LS1 s SED exhibits a strong Balmer break, which indicates that the lensed object has a relatively high surface gravity. Stars, including blue supergiants, exhibit spectra with a strong Balmer break, but stellar outbursts and explosions have low surface gravity and lack a strong Balmer break. The temperature of 11,000 14,000 K inferred from fitting the Balmer break is also substantially larger than that of almost all H-rich transients during outburst such as luminous blue variables (LBVs). While Lyman absorption of a background active galactic nucleus (AGN) at 7

8 z 9 could potentially produce a continuum break at 9500 Å, the AGN s flux blueward of the break would be almost entirely absorbed, and additional images would be expected. Our ray-tracing simulations, which are described in detail in Ref. 18, show that the MACS J1149 galaxy cluster s gravitational potential effectively increases the Einstein radii of individual stars in the intracluster medium by a factor of 100 along the line of sight to LS1. Consequently, even though intracluster stars account for 1% of the cluster s mass along the line of sight to LS1, overlapping caustics arising from intracluster stars should densely cover the source plane of the host galaxy at z = 1.49, as demonstrated by our simulation plotted in Extended Data Fig. 2. By contrast, Galactic microlensing magnification can be fully modeled using the caustic of a single star or stellar system. The ray-tracing simulations show that a star at LS1 s location should experience multiple microlensing events over a period of a decade with typical magnifications of In Fig. 4, we display the light curve of LS1 constructed from all optical and IR HST observations of the field, and we show ray-tracing simulations that can describe LS1 s light curve. 2 A Separate Microlensing Event at a Different Position A foreground gravitational lens made of smoothly distributed matter should form a pair of images of a static background source at equal angular offsets from the critical curve. However, only a single, persistent point source is apparent near the critical curve in HST imaging taken during the period We initially considered the possibility that LS1 happens to be sufficiently close 8

9 to the galaxy cluster s caustic that its pair of images have a small angular separation unresolved in HST data. As we continued to monitor the MACS J1149 cluster field, however, we detected an unexpected new source ( Lev16B ) on 30 October 2016 offset by 0.26 from LS1. We measure magnitudes of F125W = ± 0.12 AB (λ pivot = 1.25 µm) and F160W = ± 0.22 AB (λ pivot = 1.54 µm). The F125W F160W color (which corresponds approximately to rest-frame V R) of the new source is consistent with that of LS1, which has F125W F160W = 0.11 ± 0.10 mag AB. We consider that the new source could either be the counterimage of LS1, or a different lensed star. As can be seen in Fig. 1, the pair of images of LS1 s host galaxy that meet at the critical curve appear flipped relative to each other. These images are said to have opposite parity, a property of lensed images set by the sign of the determinant of the lensing magnification matrix. Assuming they are counterimages, Lev16B and LS1/Lev16A would have negative and positive parity, respectively. We have found from our ray-tracing simulations that the parity of an image of a lensed background star strongly affects its microlensing variations 18. Extended Data Fig. 2 shows that, while an image of a background star on LS1/Lev16A s side of the critical curve always has magnification of 300, its counterimage on Lev16B s side has extensive regions of much lower magnification ( 30) in the source plane. If LS1 fell in such a low-magnification region on Lev16B s side for much of the period , that could explain why LS1/Lev16B was not detected except on 30 October 2016, as shown in Extended Data Fig. 7. A 3 M object, such as a stellar binary 9

10 system, or a neutron star or black hole, can cause an image of a star to have low magnification for sufficiently long periods on Lev16B s side of the critical curve. 3 Properties of Lensed Star 1 If we assume that LS1/Lev16A and Lev16B are counterimages, then our model predicts each has an average magnification of 600. Different cluster models, however, show a factor of 2 disagreement about the magnification at LS1 s position LS1 had F125W mag in , corresponding to an absolute magnitude of M V = 9.0 ± 0.75 (sys) for a magnification of 600 per image. Post-main-sequence stars in the Small Magellanic Cloud (SMC) that have U B and B V colors approximately matching those of LS1 ( 0.40 and 0.05 mag, respectively) have luminosities reaching M V 8.8 mag 19. The two statistically significant peaks in May 2016 could correspond to a projected separation for a binary star system of 25 AU for a transverse velocity of 1000 km s 1 (see Methods). If LS1 instead consists of an unresolved pair of images, then the lensed star would need to have an offset of 0.06 pc of the caustic curve to be unresolved in HST imaging (see Methods). Its total magnification would be 10000, corresponding to a star with M V 6 mag. 10

11 4 Monte Carlo Simulation of Stellar Population Near Galaxy Cluster s Caustic We next perform simulations that allow us to estimate the probabilities (a) that LS1/Lev16A and Lev16B are counterimages of each other, and (b) of discovering a lensed star in HST galaxy-cluster observations. We use measurements of the arc underlying LS1 s position to estimate the number and luminosities of stars near the galaxy cluster s caustic. For different potential stellar luminosity functions, we calculate the number of expected bright lensed stars and microlensing events. The underlying arc extends for 0.2 ( 340 pc in the source plane) along the galaxy cluster s critical curve. If LS1/Lev16A and Lev16B are counterimages, then the lensed star is offset from the caustic by 2.2 pc in the source plane according to our lensing model. In our simulation, we populate the source plane region within 0.4 of the critical curve, or 21.9 pc from the caustic, with stars. We first need to infer the total luminosity in stars in the 21.9 pc 340 pc region adjacent to the galaxy cluster s caustic. Gravitational lensing conserves the surface brightness, and we use the arc s F125W 25 mag arcsec 2 surface brightness to compute its absolute rest-frame V surface brightness, which yields an estimate for the luminosity density of 120 L pc 2. The next step is to place stars in the 21.9 pc 340 pc region adjacent to the caustic (within 0.4 of the critical curve), whose area of 7100 pc 2 should enclose a total luminosity of L. We consider power-law luminosity functions where the number of stars with luminosity between L and L+dL is proportional to L α dl. For luminosity functions with 1.5 α 3, we normalize the 11

12 luminosity function so that the integrated luminosity equals L and compute the expected number of stars in each 0.1 L interval. We draw from a Poisson distribution to determine the number of stars in each luminosity bin, and assign each star a random position within 21.9 pc of the caustic. We next compute the average magnification µ of each star. For a lens consisting of only smooth matter, the predicted magnification at an offset R in parsecs from the caustic is µ = 880 / R. Our ray-tracing simulations find that the average magnification deviates from this prediction closer than 1.3 pc from the caustic curve (0.1 from the critical curve) due to microlensing. To estimate µ for stars closer than 0.10 to the critical curve, we interpolate in the image plane between µ = 5000 at the critical curve and µ = 680 at an offset of Our next step is to estimate the number of bright microlensing peaks (F125W 26 mag) we expect to find in existing HST observations of the MACS J1149 galaxy-cluster field. LS1 is expected to have a transverse velocity of order 1000 km s 1 relative to the cluster lens (see Methods), which corresponds approximately to LS1/Lev16A s two-week peak duration 1. If HST observations taken within a period of 10 days are counted as a single observation, then there were N obs = 50 observations of the MACS J1149 field in all optical and IR wide-band filters through 13 April 2017, and N obs = 37 observations through 15 April 2016 just before the detection of LS1. After taking into account stellar microlensing, the fraction of the source plane where the 12

13 magnification exceeds µ is (Kaiser et al., in preparation) ( f S (µ, µ) κ )( µ ) 3 ( µ ) 2, (1) where µ is the total amplification, κ is the surface density of stars making up the intracluster light (ICL) in units of the critical density, and µ is the expected magnification if the cluster consisted entirely of smoothly distributed matter. Eq. 1 does not apply at offsets smaller than 1.3 pc from the cluster caustic where the optical depth for microlensing exceeds unity. Our ray-tracing simulations indicate, however, that the formula should provide a reasonable first-order approximation at smaller distances from the caustic when we use our estimate of the average magnification µ near the critical curve 18. The number of expected microlensing events with magnification exceeding µ for each star will be N obs f S (µ, µ). The simulations provide support for the hypothesis that Lev16A and Lev16B are counterimages of LS1. Extended Data Fig. 4 shows that, if a star has an average apparent F125W brightness of at least 27.7 mag similar to LS1, then it will be responsible for 99% of F125W 26 mag events. Likewise, Extended Data Fig. 5 shows that observing a bright lensed star sufficiently close to the caustic that its images are unresolved ( 0.06 pc from the caustic) is a factor of ten less probable than observing a resolved pair of bright images of a lensed star. In nearby galaxies, the bright end of the luminosity function has a power law index of α 2.5. Young star-forming regions such as 30 Doradus, however, can have shallower functions where α 2. Extended Data Fig. 5 suggests that the probability of observing a persistent bright lensed star (F125W 27.7 mag) in the underlying arc may be 10% in existing HST observations, given 13

14 a shallow stellar luminosity function where α 2. For the steeper mean luminosity function (α 2.5) measured for nearby galaxies 20, we find a probability of %. The probability of observing at least one bright (F125W 26 mag) microlensing event is 3% for α 2, and 0.1% for α 2.5. We have repeated our simulation using the distribution of stars in 30 Doradus in the Large Magellanic Cloud (LMC), which yields similar probabilities as for the case where α 2. To estimate to first order the probability of finding a lensed star in all existing HST galaxycluster observations, we make the simplifying assumption that all strong lensing arcs have properties similar to that underlying LS1. Of the total time used to image cluster fields with HST, only at most 10% has been used to observe MACS J1149. Each of several dozen galaxy cluster fields monitored by HST contains 4 giant arcs 21. Consequently, to take into account all HST galaxycluster observations, we need to multiply our above Monte Carlo probabilities by an approximate factor of 10 4 = 40. This suggests that the probability of finding a lensed star may be reasonable, but only if the average stellar luminosity function at high redshift is shallower than α 2.5. We note that we detected a new potential source (Lev 2017A) which has a 4σ significance in the WFC3-IR imaging acquired on 3 January 2017, although the significance is only 2.5σ considering all HST imaging and the independent apertures adjacent to the critical curve. 5 Multiple Limits on the Physical Size of Lensed Star 1 Each bright microlensing peak must correspond to light from an individual star in the source plane, given the small area of high magnification adjacent to the microcaustics of intracluster stars. Ad- 14

15 ditional considerations provide evidence that the persistent source, LS1, is too compact to be a typical stellar cluster, and is instead a single stellar system (e.g., an individual star or a binary). If LS1 consists of two unresolved counterimages at the location of the critical curve, then LS1 must be more compact than 0.06 pc given the upper limit on its angular size ( ; see Methods). If Lev16A and Lev16B are instead mutual counterimages, the limit on LS1 s angular size constrains it to have a physical dimension perpendicular to the caustic of 1 2 pc, which is significantly smaller than the typical size of a stellar cluster. The absence of a persistent image at Lev16B s position places a stronger potential limit on LS1 s size. To explain the lack of a peristent counterimage at Lev16B s location, all stars in a hypothetical stellar association at LS1 s position would need to fall in a region of low magnification on the Lev16B side of the critical curve. Ray-tracing simulations indicate that LS1 would need to be smaller than 0.1 pc. A hypothetical dark matter (DM) subhalo, however, could also potentially demagnify a counterimage at Lev16B s position. 6 Inferences from LS1/Lev16A and Lev16B Assuming They Are Mutual Counterimages We next compare the HST light curve for LS1/Lev16A and Lev16B with simulated light curves created for different assumptions about intracluster stellar population and the abundance of 30 M primordial black holes (PBHs). In our ray-tracing simulations, LS1/Lev16A and Lev16B are counterimages with average magnifications from the cluster of

16 We assume (a) all intracluster light (ICL) stars are single, or (b) apply mass-dependent binary fractions and mass ratios 22. We use a stellar-mass density of M kpc 2 for a Chabrier initial mass function (IMF), or higher densities estimated in an improved analysis of M kpc 2 and M kpc 2 for Chabrier and Salpeter IMFs, respectively. The most massive star that is still living found in the intracluster medium (ICM) at z = 0.54 is assumed to have M = 1.5 M. In Extended Data Figs. 8, 9, and 10, we plot the simulated light curves for an R = 100 R lensed star where we adopt the Woosley02 23, Fryer12 24, or Spera15 25 models of stellar evolution and core-collapse physics. For steps of 50 km s 1 in the range km s 1, we stretch the simulated light curves and identify the regions that best match the data. Table 1 lists the average χ 2 value for the 150 best matches χ 2. To interpret differences in χ 2 values, we fit simulated light curves, and compute the difference χ 2 values between the χ 2 values of the generative ( true ) model and of the best-fitting model. For 68% of simulated light curves, χ 2 13, and for 95% of simulated light curves, χ For stars with 7.5 > M V > 9.5 mag, a range consistent with the most luminous stars in the SMC and LMC given the uncertainty in magnification, models constructed using a prescription for the binary fraction 22 are favored over those where all stars are single (see Extended Data Fig. 11). The χ 2 statistics also favor the Fryer12 stellar model, and a Salpeter IMF over a Chabrier IMF (see Methods). The fitting also provides evidence against models where 1% and 3% of DM consists of 30 M PBHs 26. Within the confidence intervals, the differences remain robust when extending 16

17 the upper M V limit to 10.5 mag. Table 1 also shows χ 2 values if we restrict the absolute magnitude to 7.5 < M V < 8.5 (for µ = 600), although such a low luminosity would be difficult to reconcile with LS1 s light curve. The Fryer12 model and a Salpeter IMF are still favored, but there is no preference for the binary prescription. Although our confidence intervals assume that our estimates for the stellar-mass density and magnification are correct, it may be reasonable to assume that differences in χ 2 values will be robust to modest errors in these parameters. Our cluster model also does not include DM subhalos, which could affect the average fluxes of the images. Although our fits do not favor models where 30 M PBHs account for 1% or 3% of DM, PBHs consisting of 3% of DM could produce a slowly varying average magnification, and potentially explain the absence of flux at Lev16B s position. 1. Miralda-Escude, J. The magnification of stars crossing a caustic. I - Lenses with smooth potentials. ApJ 379, (1991). 2. Kelly, P. L. et al. Multiple images of a highly magnified supernova formed by an early-type cluster galaxy lens. Science 347, (2015) Rodney, S. A. et al. SN Refsdal: Photometry and Time Delay Measurements of the First Einstein Cross Supernova. ApJ 820, 50 (2016)

18 4. Kelly, P. L. et al. Deja Vu All Over Again: The Reappearance of Supernova Refsdal. ApJ 819, L8 (2016) Oguri, M. Predicted properties of multiple images of the strongly lensed supernova SN Refsdal. MNRAS 449, L86 L89 (2015) Sharon, K. & Johnson, T. L. Revised Lens Model for the Multiply Imaged Lensed Supernova, SN Refsdal in MACS J ApJ 800, L26 (2015) Diego, J. M. et al. A free-form prediction for the reappearance of supernova Refsdal in the Hubble Frontier Fields cluster MACSJ MNRAS 456, (2016) Jauzac, M. et al. Hubble Frontier Fields: predictions for the return of SN Refsdal with the MUSE and GMOS spectrographs. MNRAS 457, (2016) Grillo, C. et al. The Story of Supernova Refsdal Told by Muse. ApJ 822, 78 (2016) Kawamata, R., Oguri, M., Ishigaki, M., Shimasaku, K. & Ouchi, M. Precise Strong Lensing Mass Modeling of Four Hubble Frontier Field Clusters and a Sample of Magnified Highredshift Galaxies. ApJ 819, 114 (2016) Treu, T. et al. Refsdal Meets Popper: Comparing Predictions of the Re-appearance of the Multiply Imaged Supernova Behind MACSJ ApJ 817, 60 (2016)

19 12. Ebeling, H. et al. A Complete Sample of 12 Very X-Ray Luminous Galaxy Clusters at z 0.5. ApJ 661, L33 L36 (2007). astro-ph/ Smith, G. P. et al. Hubble Space Telescope Observations of a Spectacular New Strong-Lensing Galaxy Cluster: MACS J at z = ApJ 707, L163 L168 (2009) Zitrin, A. & Broadhurst, T. Discovery of the Largest Known Lensed Images Formed by a Critically Convergent Lensing Cluster. ApJ 703, L132 L136 (2009) Yuan, T.-T., Kewley, L. J., Swinbank, A. M., Richard, J. & Livermore, R. C. Metallicity Gradient of a Lensed Face-on Spiral Galaxy at Redshift ApJ 732, L14 (2011) Karman, W. et al. Highly ionized region surrounding SN Refsdal revealed by MUSE. A&A 585, A27 (2016) Castelli, F. & Kurucz, R. L. New Grids of ATLAS9 Model Atmospheres. ArXiv Astrophysics e-prints (2004). astro-ph/ Diego, J. M. et al. Dark matter under the microscope: Constraining compact dark matter with caustic crossing events. ArXiv e-prints (2017) Dachs, J. Photometry of bright stars in the Small Magellanic Cloud. A&A 9, (1970). 20. Bresolin, F. et al. A Hubble Space Telescope Study of Extragalactic OB Associations. AJ 116, (1998). 19

20 21. Xu, B. et al. The Detection and Statistics of Giant Arcs behind CLASH Clusters. ApJ 817, 85 (2016) Duchêne, G. & Kraus, A. Stellar Multiplicity. ARA&A 51, (2013) Woosley, S. E., Heger, A. & Weaver, T. A. The evolution and explosion of massive stars. Reviews of Modern Physics 74, (2002). 24. Fryer, C. L. et al. Compact Remnant Mass Function: Dependence on the Explosion Mechanism and Metallicity. ApJ 749, 91 (2012) Spera, M., Mapelli, M. & Bressan, A. The mass spectrum of compact remnants from the PARSEC stellar evolution tracks. MNRAS 451, (2015) Bird, S. et al. Did LIGO Detect Dark Matter? Physical Review Letters 116, (2016) Zitrin, A. et al. Hubble Space Telescope Combined Strong and Weak Lensing Analysis of the CLASH Sample: Mass and Magnification Models and Systematic Uncertainties. ApJ 801, 44 (2015) Keeton, C. R. On modeling galaxy-scale strong lens systems. General Relativity and Gravitation 42, (2010). 29. Lotz, J. M. et al. The Frontier Fields: Survey Design and Initial Results. ApJ 837, 97 (2017)

21 30. Jones, D. O., Scolnic, D. M. & Rodney, S. A. PythonPhot: Simple DAOPHOT-type photometry in Python. Astrophysics Source Code Library (2015) Rodney, S. A. et al. Illuminating a Dark Lens : A Type Ia Supernova Magnified by the Frontier Fields Galaxy Cluster Abell ApJ 811, 70 (2015) Conroy, C., Gunn, J. E. & White, M. The Propagation of Uncertainties in Stellar Population Synthesis Modeling. I. The Relevance of Uncertain Aspects of Stellar Evolution and the Initial Mass Function to the Derived Physical Properties of Galaxies. ApJ 699, (2009) Conroy, C. & Gunn, J. E. The Propagation of Uncertainties in Stellar Population Synthesis Modeling. III. Model Calibration, Comparison, and Evaluation. ApJ 712, (2010) Kroupa, P. On the variation of the initial mass function. MNRAS 322, (2001). 35. Cardelli, J. A., Clayton, G. C. & Mathis, J. S. The relationship between infrared, optical, and ultraviolet extinction. ApJ 345, (1989). 36. Marigo, P. & Girardi, L. Evolution of asymptotic giant branch stars. I. Updated synthetic TP- AGB models and their basic calibration. A&A 469, (2007). astro-ph/ Marigo, P. et al. Evolution of asymptotic giant branch stars. II. Optical to far-infrared isochrones with improved TP-AGB models. A&A 482, (2008)

22 38. Asplund, M., Grevesse, N., Sauval, A. J. & Scott, P. The Chemical Composition of the Sun. ARA&A 47, (2009) Foreman-Mackey, D., Hogg, D. W., Lang, D. & Goodman, J. emcee: The MCMC Hammer. PASP 125, (2013) Dolphin, A. E. WFPC2 Stellar Photometry with HSTPHOT. PASP 112, (2000). astro-ph/ Watson, W. A. et al. The halo mass function through the cosmic ages. MNRAS 433, (2013) Zitrin, A. et al. New multiply-lensed galaxies identified in ACS/NIC3 observations of Cl using an improved mass model. MNRAS 396, (2009) Oguri, M. The Mass Distribution of SDSS J Revisited. PASJ 62, (2010) Jullo, E. et al. A Bayesian approach to strong lensing modelling of galaxy clusters. New Journal of Physics 9, 447 (2007) Suyu, S. H. & Halkola, A. The halos of satellite galaxies: the companion of the massive elliptical lens SL2S J A&A 524, A94 (2010) Suyu, S. H. et al. Disentangling Baryons and Dark Matter in the Spiral Gravitational Lens B ApJ 750, 10 (2012)

23 47. Schmidt, K. B. et al. Through the Looking GLASS: HST Spectroscopy of Faint Galaxies Lensed by the Frontier Fields Cluster MACSJ ApJ 782, L36 (2014) Treu, T. et al. The Grism Lens-Amplified Survey from Space (GLASS). I. Survey overview and first data release. ApJ 812, 114 (2015) Kelly, P. L. et al. SN Refsdal: Classification as a Luminous and Blue SN 1987A-like Type II Supernova. ApJ 831, 205 (2016) Chabrier, G. The Galactic Disk Mass Function: Reconciliation of the Hubble Space Telescope and Nearby Determinations. ApJ 586, L133 L136 (2003). 51. Gaudi, B. S. & Petters, A. O. Gravitational Microlensing near Caustics. I. Folds. ApJ 574, (2002). astro-ph/ Treu, T. et al. The Initial Mass Function of Early-Type Galaxies. ApJ 709, (2010) Auger, M. W. et al. Dark Matter Contraction and the Stellar Content of Massive Early-type Galaxies: Disfavoring Light Initial Mass Functions. ApJ 721, L163 L167 (2010) Spiniello, C., Koopmans, L. V. E., Trager, S. C., Czoske, O. & Treu, T. The X-Shooter Lens Survey - I. Dark matter domination and a Salpeter-type initial mass function in a massive early-type galaxy. MNRAS 417, (2011)

24 55. Cappellari, M. et al. Systematic variation of the stellar initial mass function in early-type galaxies. Nature 484, (2012) van Dokkum, P. G. & Conroy, C. A substantial population of low-mass stars in luminous elliptical galaxies. Nature 468, (2010). 57. Conroy, C. & van Dokkum, P. G. The Stellar Initial Mass Function in Early-type Galaxies From Absorption Line Spectroscopy. II. Results. ApJ 760, 71 (2012). 58. Newman, A. B., Smith, R. J., Conroy, C., Villaume, A. & van Dokkum, P. The Initial Mass Function in the Nearest Strong Lenses from SNELLS: Assessing the Consistency of Lensing, Dynamical, and Spectroscopic Constraints. ArXiv e-prints (2016) Barnabè, M. et al. A low-mass cut-off near the hydrogen burning limit for Salpeter-like initial mass functions in early-type galaxies. MNRAS 436, (2013) Conroy, C., van Dokkum, P. G. & Villaume, A. The Stellar Initial Mass Function in Early-type Galaxies from Absorption Line Spectroscopy. IV. A Super-Salpeter IMF in the Center of NGC 1407 from Non-parametric Models. ApJ 837, 166 (2017) Alves de Oliveira, C. et al. Spectroscopy of brown dwarf candidates in IC 348 and the determination of its substellar IMF down to planetary masses. A&A 549, A123 (2013) Moraux, E., Bouvier, J., Stauffer, J. R. & Cuillandre, J.-C. Brown dwarfs in the Pleiades cluster: Clues to the substellar mass function. A&A 400, (2003). astro-ph/

25 63. Renzini, A. & Ciotti, L. Transverse Dissections of the Fundamental Planes of Elliptical Galaxies and Clusters of Galaxies. ApJ 416, L49 (1993). 64. Edwards, L. O. V., Alpert, H. S., Trierweiler, I. L., Abraham, T. & Beizer, V. G. Stellar populations of BCGs, close companions and intracluster light in Abell 85, Abell 2457 and IIZw108. MNRAS 461, (2016) Smartt, S. J., Eldridge, J. J., Crockett, R. M. & Maund, J. R. The death of massive stars - I. Observational constraints on the progenitors of Type II-P supernovae. MNRAS 395, (2009) Gerke, J. R., Kochanek, C. S. & Stanek, K. Z. The search for failed supernovae with the Large Binocular Telescope: first candidates. MNRAS 450, (2015) Pejcha, O. & Thompson, T. A. The Landscape of the Neutrino Mechanism of Core-collapse Supernovae: Neutron Star and Black Hole Mass Functions, Explosion Energies, and Nickel Yields. ApJ 801, 90 (2015) Hamann, W.-R., Schoenberner, D. & Heber, U. Mass loss from extreme helium stars. A&A 116, (1982). 69. Hamann, W.-R. & Koesterke, L. Spectrum formation in clumped stellar winds: consequences for the analyses of Wolf-Rayet spectra. A&A 335, (1998). 70. Smith, N. Mass Loss: Its Effect on the Evolution and Fate of High-Mass Stars. ARA&A 52, (2014)

26 71. Belczynski, K. et al. Compact Object Modeling with the StarTrack Population Synthesis Code. ApJS 174, (2008). astro-ph/ Belczynski, K. et al. On the Maximum Mass of Stellar Black Holes. ApJ 714, (2010) Sana, H. et al. Binary Interaction Dominates the Evolution of Massive Stars. Science 337, 444 (2012). 74. Sharon, K. et al. The Type Ia Supernova Rate in Redshift Galaxy Clusters. ApJ 718, (2010) Postman, M. et al. The Cluster Lensing and Supernova Survey with Hubble: An Overview. ApJS 199, 25 (2012) Graur, O. et al. Type-Ia Supernova Rates to Redshift 2.4 from CLASH: The Cluster Lensing And Supernova Survey with Hubble. ApJ 783, 28 (2014) Freedman, W. L. The upper end of the stellar luminosity function for a sample of nearby resolved late-type galaxies. ApJ 299, (1985). 78. Grammer, S. & Humphreys, R. M. The Massive Star Population in M101. I. The Identification and Spatial Distribution of the Visually Luminous Stars. AJ 146, 114 (2013) Bastian, N. et al. Hierarchical star formation in M33: fundamental properties of the starforming regions. MNRAS 379, (2007)

27 80. Cook, D. O. et al. The connection between galaxy environment and the luminosity function slopes of star-forming regions. MNRAS 462, (2016) Oke, J. B. et al. The Keck Low-Resolution Imaging Spectrometer. PASP 107, 375 (1995). 82. Pettini, M. & Pagel, B. E. J. [OIII]/[NII] as an abundance indicator at high redshift. MN- RAS 348, L59 L63 (2004). 83. Yuan, T., Kobayashi, C. & Kewley, L. J. H II Region Metallicity Constraints Near the Site of the Strongly Lensed Supernova SN Refsdal at Redshift ApJ 804, L14 (2015) Wang, X. et al. The Grism Lens-amplified Survey from Space (GLASS). X. Sub-kiloparsec Resolution Gas-phase Metallicity Maps at Cosmic Noon behind the Hubble Frontier Fields Cluster MACS ApJ 837, 89 (2017) Steidel, C. C. et al. Strong Nebular Line Ratios in the Spectra of z 2-3 Star Forming Galaxies: First Results from KBSS-MOSFIRE. ApJ 795, 165 (2014) Maiolino, R. et al. AMAZE. I. The evolution of the mass-metallicity relation at z 3. A&A 488, (2008) Grevesse, N. & Sauval, A. J. Standard Solar Composition. Space Sci. Rev. 85, (1998). 27

28 88. Kim, A., Goobar, A. & Perlmutter, S. A Generalized K Correction for Type IA Supernovae: Comparing R-band Photometry beyond z=0.2 with B, V, and R-band Nearby Photometry. PASP 108, 190 (1996). astro-ph/ Peng, C. Y., Ho, L. C., Impey, C. D. & Rix, H.-W. Detailed Structural Decomposition of Galaxy Images. AJ 124, (2002). astro-ph/ Morishita, T. et al. Characterizing Intra-cluster light at z 0.5 in the Hubble Frontier Fields. ArXiv e-prints (2016) Schlegel, D. J., Finkbeiner, D. P. & Davis, M. Maps of Dust Infrared Emission for Use in Estimation of Reddening and Cosmic Microwave Background Radiation Foregrounds. ApJ 500, 525 (1998). 92. Kriek, M. et al. An Ultra-Deep Near-Infrared Spectrum of a Compact Quiescent Galaxy at z = 2.2. ApJ 700, (2009) Bruzual, G. & Charlot, S. Stellar population synthesis at the resolution of MNRAS 344, (2003). 94. Paczynski, B. Gravitational microlensing at large optical depth. ApJ 301, (1986). Supplementary Information is linked to the online version of the paper at Acknowledgements We express our appreciation to the Directors of the Space Telescope Science Institute, the Gemini Observatory, the GTC, and the European Southern Observatory for granting us discretionary time. We thank B. Katz, D. Kushnir, B. Periello, I. Momcheva, T. Royale, 28

29 L. Strolger, D. Coe, J. Lotz, M. L. Graham, R. Humphreys, R. Kurucz, A. Dolphin, M. Kriek, S. Rajendran, T. Davis, I. Hubeny, C. Leitherer, F. Nieva, D. Kasen, J. Mauerhan, D. Kelson, J. M. Silverman, A. Oscoz Abaz, and Z. Levay for help with the observations or other assistance. The Keck Observatory was made possible with the support of the W.M. Keck Foundation. NASA/STScI grants 14041, 14199, 14208, 14528, 14872, and provided financial support. P.L.K., A.V.F., and W.Z. are grateful for assistance from the Christopher R. Redlich Fund, the TABASGO Foundation, and the Miller Institute for Basic Research in Science (U.C. Berkeley). The work of A.V.F. was completed in part at the Aspen Center for Physics, which is supported by NSF grant PHY J.M.D. acknowledges support of projects AYA P (MINECO/FEDER, UE), AYA C02-01, and the consolider project CSD funded by the Ministerio de Economia y Competitividad. P.G.P.-G. acknowledges support from Spanish Government MINECO AYA ERC and AYA P Grants. M.O. is supported by JSPS KAKENHI Grant Numbers and 15H M.J. acknowledges support by the Science and Technology Facilities Council [grant ST/L00075X/1]. R.J.F is supported by NSF grant AST and Sloan and Packard Foundation Fellowships. M.N. acknowledges support from PRIN-INAF O.G. was supported by NSF Fellowship under award AST J.H. acknowledges support from a VILLUM FONDEN Investigator Grant [number 16599]. HST imaging was obtained at Author Contributions P.K. planned and analyzed observations, wrote the manuscript, and developed simulations. P.K., S.R., P.G.P-G., T.M., M.J., J.S., A.V.F., J.H., M.G., D.H., B.L.F., M.B., W.Z., G.B, A.M.B., A.C., L.C., C.G., J-P.K., T.M., C.M., M.N., I.P.-F., A.G.R., P.R., K.B.S, and 29

30 B.J.W. obtained follow-up HST and ground-based imaging. J.M.D. developed microlensing simulations. S.R., T.B., A.Z., T.T., P.G.P-G., M.J., M.O., A.V.F., N.S., J.H., B.L.F., and S.E.d.M. helped to prepare the manuscript. N.K. interpreted the microlensing events and derived analytic rate formula. T.B., A.Z., T.T., M.J., M.O., X.W., S.W.J., R.J.F., S.E.d.M., O.G., and B.J.W. aided the interpretation. P.G.P-G. modeled the arc s SED. T.M. modeled the ICL. M.O., J.R., R.K., and K.S. modeled the galaxy cluster. L.P., and C.N. considered the possibility that Icarus could exhibit diffraction effects. T.W.R. analyzed the microlensing simulations. N.S. aided interpretation of the star s SED. X.W. estimated the gas-phase metallicity at Icarus location. M.N. extracted photometry of Icarus using a complementary pipeline. Author Information Data used for this publication may be retrieved from the NASA Mikulski Archive for Space Telescopes ( Reprints and permissions information is available at The authors declare no competing financial interests. Correspondence should be addressed to P.K. (patrick.l.kelly@gmail.com). 30

31 a b Images of host galaxy (z = 1.49) 4ʺ 2014 Stellar Images: LS1 / Lev 2016A + Lev 2016B (+ Lev 2017A) c 2ʺ SN Refsdal LS1 / Lev 2016A Late May Figure 1: Locations of lensing events coinciding with background spiral galaxy near the MACS J1149 galaxy cluster s critical curve. Left panel shows the positions of magnified images of stars LS1 / Lev16A and Lev16B as well as candidate Lev 2017A close to ( ) the critical curve, where magnification rises rapidly. Dashed line shows the location of the critical curve from the CATS cluster model 8. The Einstein Cross formed from yellow point sources consists of images of SN Refsdal 2. Right panel shows the field in 2014, and in May 2016 when LS1 exhibited a microlensing peak. 31

32 1ʺ LS1/Lev16A Figure 2: Proximity of LS1/Lev16A to the MACS J1149 galaxy-cluster critical curve for multiple galaxy-cluster lens models. Critical curves for models with available high-resolution lens maps including Ref. 8 (CATS; solid red), Ref. 27 (short-dash orange), Ref. 10 (solid blue), and Ref. 28 (long-dash black) are superposed on the HST WFC3-IR F125W image. Although predictions for the location of the critical curve near LS1 disagree by 0.25, LS1 lies within 0.13 of all of these models predictions. 32

33 Rest Wavelength (Å) Flux (µjy) Underlying Point Source Flux Excess May 2016 (rescaled; χ 2 ν=1.5) Observed Wavelength (Å) Magnitude (AB) Figure 3: The SEDs of LS1 measured in (red) and of the rescaled, excess flux density at LS1 s position close to its May 2016 peak (Lev16A; black) are consistent. Rescaling the SED of the flux excess to match to that of the source yields χ 2 ν = 1.5, indicating that they are statistically consistent with each other despite a flux density difference of a factor of 4. The SED shows a strong Balmer break consistent with the host-galaxy redshift of 1.49, and stellar atmosphere models 17 of a mid-to-late B-type star provide a reasonable fit. The blue curve has T = 11, 180 K, log g = 2, A V = 0.02, and χ 2 = 16.3; the orange curve has T = 12, 250 K, log g = 4, A V = 0.08, and χ 2 = 30.6; the black curve has T = 12, 375 K, log g = 2, A V = 0.08, and χ 2 = 12.9; and the green curve has T = 13, 591 K, log g = 4, A V = 0.13, and χ 2 = Black circles show the expected flux density for each model. 33

34 Flux (µjy) Spera15 ( 2 =184.6) Spera15/3% PBH ( 2 =219.6) Date (MJD) Near Peak Magnitude (AB) Flux (µjy) Near Peak (May 2016) Date (MJD) Magnitude (AB) Figure 4: Light curve of the magnified star LS1, and best-matching simulated light curves during each interval. Fluxes measured through all wide-band HST filters are converted to F125W Fig. 3: Light curve of the magnified star LS1, and best-matching simulated light curves during each interval. Fluxes measured through all wide-band HST filters are converted to using LS1 s SED. The upper panel shows LS1 s full HST light curve which begins in The F 125W using LS1 s SED. Upper panel shows LS1 s full HST light curve which begins in lower panel shows the most densely sampled part of the light curve including the May 2016 peak (Lev16A). The lower This maximum panel shows shows the two most successive densely peaks sampled that part may of correspond the light curve to a lensed including binary the system May of 2016 stars peak at z = (Lev A). This maximum shows two successive peaks that may correspond to a lensed binary system of stars at redshift z =

35 2014 October January Lev 2017A Lev 2016B LS1 / Lev 2016A LS1 / Lev 2016A 0.5 LS1 / Lev 2016A Figure 5: Highly magnified stellar images located near the MACS J1149 galaxy cluster s critical curve. The left panel shows LS1 in 2014; we detected LS1 when it temporarily brightened by a factor of 4 in late-april The center panel shows the appearance of a new image dubbed Lev16B on 30 October The solid red curve marks the location of the cluster s critical curve from the CATS cluster model 8, and the dashed lines show the approximate 1σ uncertainty from comparison of multiple cluster lens models The position is consistent with the possibility that it is a counterimage of LS1. The right panel shows the candidate named Lev 2017A with 4σ significance detected on 3 January If a microlensing peak, Lev 2017A must correspond to a different star. 35

36 7.50 > M V > > M V > Best Matches (406 yr) 150 Best Matches (406 yr) χ 2 Σ Model PBH T IMF χ 2 Σ Model PBH T IMF Low Stellar-Mass Density Best L Fryer12 B Cha L Fryer12 B Cha L Woosley02 B Cha L Spera15 S Cha L Spera15 B Cha L Woosley02 S Cha L Woosley02 S Cha L Spera15 3% S Cha L Spera15 S Cha L Woosley02 B Cha L Fryer12 S Cha L Spera15 B Cha L Spera15 1% S Cha L Fryer12 S Cha Worst L Spera15 3% S Cha L Spera15 1% S Cha High Stellar-Mass Density Best H Spera15 B Sal H Spera15 B Sal Worst H Spera15 B Cha H Spera15 B Cha Table 1: Comparison between model light curves and observed light curves for LS1 / Lev16A and Lev16B. Measured χ 2 statistics provide evidence about the binary fraction and IMF of the stellar population responsible for the intracluster light (ICL), distinguish among stellar evolution and supernova models 23 25, and disfavor the possibility that 1 3% of dark matter consists of 30 M PBHs. To interpret differences in χ 2 values, we fit simulated light curves, and compute the difference χ 2 values between the χ 2 values of the generative ( true ) model and of the best-fitting model. For 68% of simulated light curves, χ 2 13, and for 95% of simulated light curves, χ These simulations assume that our estimates for the galaxy cluster s magnification and stellar-mass density are correct. The Σ column denotes whether light curves were computed with a low (L) or high (H) stellar-mass density. The Type column specified whether the stars are single (S), or have the mass-dependent binary fraction and mass ratios of stars determined at low redshift (B) 22. The IMF column specifies whether a Chabrier ( Cha ) or Salpeter ( Sal ) IMF was used to assemble the ICL stellar population. 36

37 Methods: HST Imaging. The HST observations include those from GO programs (and Principal Investigators) (M. Postman), (T.T.), (J. Lotz), and (S.R.), and 14041, 14199, 14528, 14872, and (P.K.). Constructing Light Curve. All optical and IR HST imaging of the MACS J1149 field with moderate depth has yielded a detection of LS1. For each instrument and wide-band filter combination, we constructed a light curve for LS1. We first measured LS1 s flux in a deep template coaddition of Hubble Frontier Fields and early SN Refsdal follow-up imaging. The Hubble Frontier Fields program 29 acquired deep imaging of the MACS J1149 galaxy cluster between November 2013 and May 2015 in the ACS WFC F435W (λ c = 0.43 µm), F606W (λ c = 0.59 µm), and F814W (λ c = 0.81 µm), and the WFC3 IR F105W (λ pivot = 1.05 µm), F125W (λ pivot = 1.25 µm), F140W (λ pivot = 1.39 µm), and F160W (λ pivot = 1.54 µm) wide-band filters. The second step was to measure the differences in LS1 s flux between the deep template coadded image and at each imaging epoch. We accomplished this latter step by subtracting the deep template coaddition from coadditions of imaging at each epoch, and measuring the change in LS1 s flux from these resulting difference images. To measure LS1 s flux in each deep template coaddition, we first fit and then subtracted the ICL surrounding the BCG. We next measured the flux at LS1 s position inside an aperture radius of r = 0.10 using the PythonPhot package 30. To measure the uncertainty in the background from 37

38 the underlying arc, we placed a series of four nonoverlapping apertures having r = 0.10 along it. We use the standard deviation of these aperture fluxes as an estimate of the uncertainty in the background. Aperture corrections were calculated from coadditions of the standard stars P330E and G191B2B 31. For each HST visit, we created a coadded image of all exposures acquired in each wideband filter. We next subtracted the deep template coaddition from the visit coaddition to create a difference image. Using the PythonPhot package 30, we measured the flux inside of an r = 0.10 circular aperture in the difference image. We finally computed the total flux at each epoch by adding the flux measured from the deep template coaddition and that measured from each difference image. There is no source apparent at Lev16B s position in deep template coadditions. Therefore, we do not add any flux measured from the deep template coaddition to the light curve we construct for Lev16B, which is plotted in Extended Data Fig. 7. Estimating LS1 s Color. LS1 s brightness changed between the epochs when the deep template imaging was acquired by the Hubble Frontier Fields and SN Refsdal follow-up programs. However, the MACS J1149 cluster field was monitored using F125W (and F160W) with a cadence of 2 weeks after the discovery of SN Refsdal in November We measured LS1 s F125W light curve from these data, and performed a fit to the light curve using using a third-order polynomial. 38

39 We used the polynomial to estimate LS1 s F125W flux at the average epoch when the deep template imaging in each detector and filter was acquired. The ratio between the F125W flux and that in the other filter (e.g., F140W) provides a measurement of LS1 s color (F140W F125W, in this example). We restricted the Hubble Frontier Fields imaging to that taken between November 2014 and May 2015, since monitoring of SN Refsdal is available to construct LS1 s IR light curve beginning in November Creating a Combined Light Curve. We used our estimates of LS1 s color to convert the light curves measured in all available optical and IR filters to F125W light curves, and combined them. We also binned all F125W observations to construct the combined light curve plotted in Fig. 4 and used to fit models. Constraints on the Age of Stellar Population in Underlying Arc To compute models using Flexible Stellar Population Synthesis (FSPS) 32,33, we adopt a simple stellar population with an instantaneous burst of star formation, and include nebular and continuum emission. We use Python bindings ( to FSPS to calculate simple stellar populations with a Kroupa IMF 34 and a Cardelli extinction law 35 with R V = 3.1. These models use the Padova isochrones 36,37. A recent analysis finds a solar oxygen abundance of 12 + log(o/h) = 8.69 ± 0.05 dex and a solar metallicity 38 of Z = We therefore calculate models using log(z/z ) = 0.35, which 39

40 corresponds to Z = and is best matched with the FSPS parameter zmet = 15. To estimate the age and dust extinction of the adjacent stellar population along the arc, we use emcee 39, which is an implementation of a Markov Chain Monte Carlo (MCMC) ensemble sampler. We adopt a uniform prior on the stellar population age from 0 to 3 Gyr, and a uniform prior on the extinction A V from 0 to 2 mag. Ability to Detect Pair of Images of a Star Adjacent to Cluster Caustic. A possibility is that we do not observe a pair of images of LS1, because it is very close to the cluster caustic and the available HST imaging is not able to resolve its two images. Here we calculate how close the star must be to the caustic. If LS1 lay very close to the critical curve, then Lev16B would correpond to the microlensing event of a different star, at an offset of 0.26 from the cluster critical curve. For a star sufficiently close to the caustic, the pair of images will not be resolved with HST. Our simulations suggest that demagnifying one of the two images will be unlikely when the star is close to the cluster caustic, although they do not include expected dark-matter subhalos. The angular resolution of HST is greatest in the F606W band (λ c = 0.59 µm) and almost as sharp in the F814W band (λ c = 0.81 µm), and observations in these wide-band filters provide the best opportunity to test whether LS1 consists of two adjacent images. We use coadditions of imaging taken by the Hubble Frontier Fields program. To determine the limit we can place on the separation of two possible images, we inject pairs of point sources having the same combined magnitude as the images made using the ACS WFC F606W and ACS WFC F814W exposures at each epoch. 40

41 The full width at half-maximum intensity (FWHM) of our ACS point-spread-function (PSF) models constructed from observations of the standard stars P330E and G191B2B only agree within 10% with the measured FWHM of the stars in our coadded images. Consequently, the measured FWHM of the injected pairs of PSFs should not correspond directly to what we would we measure in the ACS data. Therefore, we compute the fractional increase in the measured FWHM with the increasing separation between the pair of injected PSFs. Next, we multiply the FWHM estimated from the stars in each image by this factor to compute limits from the ACS imaging. We inject 100 fake stars with the same magnitude using models of the ACS WFC F606W and F814W PSF. After injecting the point sources in a grid, we use the IRAF task imexam to estimate the Gaussian FWHM (GFWHM) using the comma command from the simulated data. The imexam model we use has a Gaussian profile, and we specify three radius adjustments while the fit is optimized. Pixels are fit within 3 pix (0.03 pix 1 ) of the center, and we use a background buffer of 5 pix. These simulations show that any separation between the two images greater than can be detected >95% of the time. The upper limit implies that the star would have to be closer than 0.06 pc to the cluster caustic. As we show in Extended Data Fig. 5, the relative probability of a persistently bright (F125W < 27.7 mag) star being located within 0.06 pc is 10%. DOLPHOT 40 fit parameters for the images of LS1, Lev16B, and Lev 2017A fall in the range expected for point sources in the DOLPHOT reference manual com/dolphot/dolphot.pdf, although these criteria are not highly sensitive to a pair of images. 41

42 Transverse Velocity of Star. We use the following expression [Equation 12 of Ref. 1] for the apparent transverse velocity of a lensed source: v = v s v o (1 + z s ) D s(v l v 0 ) D l (1 + z l ), (2) where D l and D s are the angular-diameter distances of the lens and source, and v 0, v l, and v s are (respectively) the transverse velocities of the observer, the lens, and the source with respect to the caustic. The expression only applies for a universe without spatial curvature. Cosmological simulations have found that merging galaxy-cluster halos and subhalos have pairwise velocities of km s 1 with tails to lower and higher velocity 41. Given the expected velocity of the lens, the peculiar velocities of Earth ( 400 km s 1 ) and of the host galaxy relative to the Hubble flow, and the motion of the star (< km s 1 ) relative to its host galaxy, a typical transverse velocity should be 1000 km s 1. In our light-curve fitting analysis, we consider transverse velocities of km s 1. Intrinsic Luminosity of Lensed Star. While extremely luminous stars are rare in the nearby universe, they require smaller magnification and can be at a greater distance from the caustic. For a lens with a smooth distribution of matter, the magnification µ falls within the distance d from the caustic as µ 1/ d. Therefore, the area A in the source plane in which the magnification is greater than µ scales as A(> µ) 1/µ 2. The observed flux F of a lensed object is F Lµ, where the object s luminosity is L. Therefore, a star with luminosity L appears brighter than f inside an area A(> f ; L) L 2. 42

43 Galaxy-Cluster Lens Model. Prior to the identification of LS1 in late-april 2016, the cluster potential had been modeled using the codes LTM 27,42, WSLAP+ 7, GLAFIC 5,10,43, LENSTOOL 6,8,44, and GLEE 9,45,46. These used several different sets of multiply imaged galaxies 11, which included new data from the Grism-Lensed Survey from Space (GLASS; PI Treu) 47,48, MUSE (PI Grillo) 9, and grism follow-up observations of SN Refsdal (PI Kelly) 49. In Fig. 5, we plot as an example the position of the critical curve from the CATS model 8 created using LENSTOOL 44, showing that it passes close to the position of LS1. The critical curves of all of these models, however, pass within similarly small offsets from LS1 s coordinates. For the simulation of the light curves of a star passing close to the cluster caustic, we use the WSLAP+ model of the cluster mass distribution 7 and draw stars randomly from a Chabrier IMF 50 until the stellar mass density equals the value we estimate of M kpc 2 for a Chabrier IMF and M kpc 2 for a Salpeter IMF. The WSLAP+ cluster lens model, which includes only smoothly distributed matter and cluster galaxies, yields several important relations describing the magnification near the critical curve, and the relationship between lensed θ and unlensed β angles. The magnification µ for a smooth cluster model [e.g., Eq. 35 of Ref. 51] can be described as µ = 155/θ, where θ is the observed angular offset from the critical curve in arcseconds, and µ = 19/ β relates the unlensed angular position β and µ. The angles β and θ follow the relation β = θ 2 /66.5, and both β and θ are in units of arcseconds. 43

44 We simulate the light curves of caustic-crossing events using a resolution of 1 µarcsec per pixel over an area of 83 pc 6.5 pc in the lens plane. This area is aligned in the direction where a background source moving toward the cluster caustic would appear to be moving. If the background star is moving with an apparent velocity of 1000 km s 1 in the source plane, its associated counterimage would take 400 yr to cross the 83 pc of the simulated region, which corresponds to ( ) yr 1. The lensed star is given a transverse velocity of 1000 km s 1 in the source plane; the resulting light curve can be streched to simulate different transverse velocities. Owing to the high magnification, the counterimages apparent motion in the image plane is large. N-body simulations show that clusters of galaxies contain (and are surrounded by) a large number of subhalos. Smaller subhalos near the cluster center may not survive the tidal forces of the cluster and are easily disrupted. The larger surviving halos and smaller subhalos along the line of sight can produce small distortions in the deflection field that could in principle distort the critical curve (and caustic). Lens models of MACS J1149 do predict such distortions around the member galaxies. However, since the typical scales of the distortion in the deflection field are proportional to the square root of the mass of the lens, the distortions from the surviving halos are orders of magnitude larger than the scale of the distortion associated to the microlenses (from the ICL). Consequently, on the scales relevant for this work ( 0.2 ), the combined deflection field of the cluster plus the DM substructure can still be considered as a smooth distribution and the critical curve could still be well approximated by a straight line. For a cluster model populated with stars in the ICM, Extended Data Fig. 14 shows the trains 44

45 or multiple counterimages of a single background star near the cluster s caustic. Replacing the cluster s smoothly varying matter distribution with an increasing fraction of 30 M PBHs yields an increasingly long train, although its expected extent ( 3 milliarcsec) when PBHs account for 10% of DM would be smaller than would be possible to detect in the HST imaging. Initial Mass Function for Stellar and Substellar Objects. Strong lensing and kinematic 52 55, as well as spectroscopic 56,57 analyses of early-type galaxies have found evidence that the IMF of stars in early-type galaxies may be bottom-heavy a larger fraction of stars have subsolar masses than is observed in the Milky Way. Spectroscopic evidence for a Salpeter-like bottom-heavy IMF in the inner regions of early-type galaxies comes from the strength of spectral features sensitive to the surface gravity of stars with M 0.3 M 56,57. However, these two sets of diagnostics do not always show agreement in the same galaxies, and the discrepancy is not yet understood 58. In Extended Data Fig. 9, we show that a Salpeter IMF yields a substantially higher frequency of microlensing peaks than a Chabrier IMF. In stellar kinematics and strong lensing, the DM is assumed to follow a simple parameteric (e.g., power-law) function near the galaxy center, while stellar matter is assumed to trace the optical emission. The total matter profile inferred from observations is decomposed into stellar and DM components, and the M /L ratio of the stellar component is used to place constraints on the IMF. Substellar objects having masses below the H-burning limit (M 0.08 M ) are not generally included as a component of the stellar mass in kinematic and lensing analyses. Substellar masses, 45

46 however, should also trace the stellar mass distribution. The inferred M /L ratios near the centers of elliptical galaxies are approximately twice as large as those expected for Milky-Way-like Chabrier IMFs, e.g., Refs If the stellar IMF in early-type galaxies has a Salpeter slope, the ratio of 2 would imply that a Salpeter IMF cannot extend to object masses significantly smaller than the H-burning limit 59. Indeed, the integral of the Salpeter IMF from zero mass through the H-burning limit diverges, so the IMF of substellar objects must be less steep than Salpeter below the 0.08 M. The integral of a Chabrier IMF in the range 0 < M < 0.10 M is 10% of the integral in 0.10 < M < 100 M. High signal-to-noise-ratio spectra of NGC 1407 are best fit by a super-salpeter IMF (Γ = 1.7; dn/d log m m Γ ) to the H-burning limit 60. In the Milky Way, surveys of substellar objects find that their mass function is likely flat or declining with decreasing mass. IR imaging of the young Milky-Way cluster IC 348 yields a population of brown dwarf stars consistent with log-normal mass distribution 61. Γ = 0.0 and Γ = 0.3 provide a reasonable fit to the populations of objects with masses smaller than 0.1 M in IC 348 and Rho Oph, respectively. Analysis of the Pleiades open clusters to 0.03 M found a population consistent with a log-normal distribution with m c = 0.25 M and σ log m = For this analysis, we only include objects with initial masses greater than 0.01 M. For the light curves generated with a Chabrier IMF, we assume that the IMF continues to this lower-mass cutoff. For the Salpeter light curves, the Salpeter form truncates at 0.05 M ; for lower initial masses, we assume that the number density of objects is constant in logarithmic intervals. 46

47 Mass Function of Surviving Stars and Compact Remnants in the ICM. For a given a star-formation history, GALAXEV computes the mass in surviving stars and in remnants using the Renzini93 prescription for the mapping between zero-age main-sequence masses and remnant masses (the initial final mass function) 63. Dead stars with initial masses M i < 8.5 M become white dwarfs with mass M M i ; those with 8.5 M M i < 40 M become 1.4 M neutron stars; and those with M i 40 M become BHs with 0.5 M i. We assume that the most massive surviving star found in the ICM at z = 0.54 has a mass of 1.5 M, approximately the expected value for a 4.5 Gyr stellar population. For stars with masses 1.5 M, we use three separate theoretical initial final mass functions to compute the distribution of remnant masses. The evolution of massive stars and the mass of their remnants is expected to depend on the stars mass-loss rate, which is thought to vary significantly with their metallicity. Integral field-unit (IFU) spectroscopy of low-redshift galaxy clusters has been able to place approximate constraints on the metallicity and age of the stars found in the ICM. IFU spectroscopy within 75 kpc of the BCGs of the nearby Abell 85, Abell 2457, and II Zw 108 galaxy clusters found that the ICL light can be best fit by a combination of substantial contributions from an old population ( 13 Gyr) with high metallicity (Z 2 Z ) and from a younger population ( 5 Gyr) with low metallicity (Z 0.5 Z )

48 Light-Curve Fitting. To fit the LS1 / Lev16A and Lev16B light curves, we identify the peaks in the simulated light curves that are 2σ above each light curve s mean magnification. We next stretch the model light curves in time for transverse velocities in the range km s 1 in steps of 50 km s 1. For each light curve and velocity, we find a best-fitting solution for a series of intervals in absolute magnitudes between M V = 7 and M V = 10.5 mag in increments of 0.5 mag. For each interval and peak, we find the best-fitting value of M V within the upper and lower bounds in luminosity. For the set of fits at each transverse velocity and each M V interval, we rank all peaks according to the χ 2 values separately for LS1 / Lev16A and Lev16B. We then pair these ranked lists of best-fitting peaks (i.e., the best-fitting peak for LS1 / Lev16A is matched with that for Lev16B, etc.), and add the χ 2 values for each pair together. Next, we identify the best χ 2 values for all values of transverse velocity and ranges in absolute luminosity, and assemble a list of these best χ 2 values. Our goodness-of-fit statistic χ 2 is the average of the 150 best χ 2 values. Interpreting the χ 2 Statistic Using Simulated Light Curves. To interpret the χ 2 values, we generate fake light curves for each of the models listed in Table 1. The simulation for each model yields a magnification over a 406 yr period for a transverse velocity of 1000 km s 1. For lensed stars with absolute magnitudes M V of 8, 9, and 10, we create simulated apparent light curves, and we append the light curve after reversing the temporal axis to create effectively an 812 yr light curve. 48

49 For each simulated light curve, we identify all peaks where the apparent F125W AB magnitude is brighter than 26.5 mag. For each peak, we randomly select a transverse velocity drawn from a uniform distribution in the interval km s 1. We use the cadence and flux uncertainties of the measured light curve of LS1 to generate a fake light curve. We next shift the peak of the measured light curve of LS1 / Lev16A (or Lev16B) to match the peak of the simulated model light curve. We then create a fake observation by sampling the simulated light curve at the same epochs as the actual measurements, and adding Gaussian noise matching the measurement uncertainties. For each such simulated light curve, we compute the χ 2 statistic using the full set of models. The region of the simulated light curve used to created the fake data set is excluded from fitting. As shown in Extended Data Fig. 13, the combined χ 2 statistic we measure for LS1 / Lev16A and Lev16B falls inside of the expected range of values. We note that a significant fraction of simulated light curves have average values of χ 2 >1000, implying that they are not well-fit by other regions of the simulated light curves. For all simulated light curve where the average χ 2 value is within 100 of the value we measure for LS1 / Lev16A and Lev16B, we calculate the difference χ 2 values between the χ 2 values of the generative ( true ) model and of the best-fitting model. For 68% of simulated light curves, χ 2 13, and, for 95% of simulated light curves, χ Massive Stellar Evolution Models The fates of massive stars remain poorly understood owing to the complexity of massive stellar evolution and the physics of supernova (SN) explo- 49

50 sions. Indirect evidence suggests that a fraction of massive stars may collapse directly to a black hole instead of exploding successfully 65,66, due to (for example) failure of the neutrino mechanism 67. We compute light curves and magnification maps using three sets of predictions for the initial final mass function As a first model, we adopt the initial final mass function predicted by the solar-metallicity, single stellar evolution models 23 (Woosley02) [Fig. 9 of Ref. 24]. In the Woosley02 models, the prescription for driven mass-loss rate at solar metallicity causes stars with initial masses 33 M to end their lives with significantly reduced He core masses, leading such stars to become BH remnants with masses no larger than 5 10 M. The Woosley02 mass-loss prescription uses theoretical models of radiation-driven winds for OB-type stars with T > 15, 000 K, and empirical estimates for Wolf-Rayet stars 68 that have been adjusted downward by a factor of three to account for the effects of clumping in the stellar wind 69. The mass-loss rate for single O- type stars during their main-sequence evolution may have been overestimated by a factor of 2 3, owing to unmodeled clumping in their winds 70. A second (Fryer12) initial final mass function was computed for single stars at subsolar metallicity (Z = 0.3 Z ; Z = 0.006) [the DELAYED curve in Fig. 11 of Ref. 24]. These predictions use the StarTrack population synthesis code 71,72. According to this model, BHs with masses up to 30 M form from the collapse of massive stars. Finally, we use a third initial final mass function (Spera15) [Fig. 6 of Ref. 25] to calculate the masses of remnants for the stellar population making up the ICM. The Spera15 relation we adopt was computed using the PARSEC evolution tracks for stars with metallicity Z = 0.006, and 50

51 explosion models where the SN is delayed, occurring 0.5 s after the initial bounce. According to the Spera15 initial final mass relation we adopt, stars having initial masses 33 M become BHs with masses within the range M. Approximately 70% of massive stars exchange mass with a companion, while 1/3 of stars will merge 73. It is also possible that the success of explosions is not related in a simple way to the stars initial mass or density structure, given the potentially complex dependence of the critical neutrino luminosity for a successful explosion on these progenitor properties 67. Observations of Similar Cluster Fields with HST. Massive galaxy clusters have been the target of extensive HST imaging and grism-spectroscopy campaigns in the last several years, and these need to be taken into account when considering the probability of finding a highly magnified star microlensed by stars making up the ICL. Detecting transients requires at least two separate observing epochs, which is possible only for programs designed with more than a single visit, or when archival imaging is available. In addition to smaller search efforts 74, large programs have been the Cluster Lensing and Supernova survey with Hubble (CLASH) 75, the GLASS 47,48, the Hubble Frontier Fields 29, and the Reionization Lensing Cluster Survey (RELICS). Transients in Hubble Frontier Fields and GLASS imaging have been subsequently observed by the FrontierSN program. With a total of 524 orbits, CLASH acquired imaging of 25 galaxy-cluster fields. A search for transients in the CLASH imaging made use of template archival imaging when available. The survey acquired imaging over a period of three months of each cluster field with repeated visits in 51

52 each filter. Each epoch had an integration time of s. The systematic search for transients in CLASH imaging had near-infrared limiting magnitudes at each epoch of F125W 26.6 and F160W 26.7 mag AB [see Tab. 1 of Ref. 76]. The Hubble Frontier Fields program used 140 orbits to observe each of 6 galaxy-cluster fields (total of 840 orbits). For each cluster, ACS optical and WFC3 IR imaging split into in two campaigns, each of which lasted for approximately a month. These were separated from each other by a period of 6 months to allow the telescope roll angles to differ by 180, to image a parallel field adjacent to the cluster with the same instruments. Each of six to twelve epochs in each WFC3 IR wideband filter for each cluster field had an integration time of 5500 s. The systematic search for transients in near-infrared Hubble Frontier Field imaging by the FrontierSN team (PI: Rodney) had near-infrared limiting magnitudes at each epoch of F125W 27.5 and F160W 27.2 mag AB. The GLASS survey 47,48 acquired WFC3 IR grism spectroscopy of ten galaxy-cluster fields over 140 orbits. The survey acquired direct pre-imaging in the WFC3 IR F105W and F140W over four epochs The probability of observing a luminous star adjacent to a caustic will depend on the number of lensed galaxies that overlap a galaxy cluster s caustic. For massive clusters, caustic curves coincide with the ICL, so magnified stars should exhibit microlensing fluctuations. Earlier work has estimated that a 10 5 L star (M V 7.5 mag) in a giant arc (with z = 0.7) and crossing the caustic of cluster Abell 370 (z = 0.375), a Hubble Frontier Fields target, would remain brighter 52

53 than V = 28 mag for 700 yr, given a V-band limiting magnitude of 28 and a 300 km s 1 transverse velocity 1. A study of HST imaging of CLASH galaxy-cluster fields 21 has found that each cluster field contains 4 ± 1 giant arcs with length 6 and length-to-width ratio 7. Given that the host galaxy of LS1 would not be classified as a giant arc according to these criteria, it is likely that additional galaxies in each field may lie on the cluster caustic. An approximate census finds that 10% of HST galaxy-cluster observing time has been used to image MACS J1149. Considering the full set of HST cluster observations, and the results of our Monte Carlo simulations for the single arc underlying LS1 (1 3% for a shallow α 2 luminosity function; % for α 2.5), the probability of observing at least one bright magnified star adjacent to a critical curve should be appreciable, in particular if the average luminosity function is more shallow at high redshifts than in the nearby universe. 53

54 Supplementary Information: Simulation Using Stars Near Center of 30 Doradus. We used the SIMBAD ( catalog to retrieve information for objects in the H ii region 30 Doradus in the LMC. Stars were first selected from the catalog having V-band magnitudes and closer than 20 to its center ( 280 pc). We include objects classified as stars, and exclude F,G,K, or M type stars unless they are classified as supergiants. Using a distance of 49 kpc to the LMC, we calculated the stars absolute magnitudes M V, and their distances from the center of the cluster. For 10,000 trials, we randomly placed the cluster s CC within 20 pc of the 30 Doradus center, and rotated the stars around the center of 30 Doradus by an angle drawn from a uniform distribution. We find a probability of 1% of finding a star with a persistent average brightness of at least 27.7 mag, and we find that such a star will also be responsible for 99% of <26 mag microlensing events. These probabilities are similar to those we estimated from our simulation where the positions of luminous stars near the CC were drawn randomly from a uniform distribution Slope of Stellar Luminosity Function in 30 Doradus. We have also used the SIMBAD catalog of stellar sources to estimate the stellar luminosity function of bright stars in 30 Doradus. Placing stars in 0.5 mag bins by their absolute V-band magnitudes, we measure a power-law index of α 2. No correction is applied for crowding, or the binary fraction. 54

55 Slope of the Stellar Luminosity Function in Nearby Galaxies. Stars found in OB associations in seven nearby galaxies observed with HST show a luminosity function of α = 2.53± The stellar luminosity function for stars more luminous than M V < 8.5 mag is not well constrained, the number counts of the M V < 8.5 mag stars in this study are consistent with the slope measured for stars with 8.5 < M V < 5 mag. The slope of α = 2.53 ± 0.08 agrees approximately with a separate earlier analysis 77, which studied the slope of the upper end of the stellar luminosity function of the bluest stars in nearby galaxies using ground-based imaging and found α = 2.68 ± 0.08, although the latter analysis extended only to M V 9.5 mag [see Fig. 7 of Ref. 77]. A second census of the stellar population in galaxy M101 shows that it may host a small number of luminous stars with absolute magnitude 10 M V 11 [see Fig. 7 of Ref. 78]. The luminosity function of OB associations 79 can be well described by a power-law function having an index α 2. The luminosity function of star-forming regions may become flatter in galaxies with higher star-formation rates and star-formation rate densities 80. Ground-Based Follow-up Campaigns. We observed the field with direct imaging with the Low Resolution Imaging Spectrometer (LRIS) 81 on the Keck-I 10 m telescope on 6 May 2016 (PI Filippenko). Director s Discretionary programs with the GTC (PI Pérez González; GTC ), the Very Large Telescope (PI Selsing; 297.A-5026), Gemini North (PI Kelly; GN-2016A- DD-8), and the Discovery Channel Telescope (PI Cenko) obtained follow-up imaging in optical bandpasses. 55

56 Detection from the Ground with the Gran Telescopio Canarias. We obtained i -band observations of the MACS J1149 field with the 10.4 m Gran Telescopio Canarias (GTC) on 6 June 2016 and 7 June 2016, after the May 2016 peak. To estimate the flux at LS1 s position, we extracted the flux inside several apertures with diameters within 1 2 times the the PSF FWHM, and applied aperture corrections to obtain integrated fluxes. The flux estimates for the different apertures agree within mag. The i -band AB magnitudes are ± 0.52 on MJD in conditions with 1.0 seeing and a total integration of 3000 s, and ± 0.43 on MJD in 0.8 seeing and a total integration of 5430 s. Metallicity of the Local Host-Galaxy Environment of LS1. The gas-phase oxygen abundance of LS1 s host galaxy, including that within the immediate environment of SN Refsdal, has been studied using multiple datasets. LS1 and SN Refsdal have similar offsets from the host nucleus (within 0.5 kpc), so the local metallicity near LS1 s and SN Refsdal s host-galaxy locations should have similar values. Both the CATS model 8 ( 6.7 kpc and 7.3 kpc, respectively) and the GLAFIC model 10 ( 7.9 kpc and 8.2 kpc) find similar nuclear offsets for LS1 and SN Refsdal. Analysis of Keck-II OSIRIS integral-field unit (IFU) spectra reported a 3σ upper limit of 12 + log(o/h) < 8.67 dex in the Pettini & Pagel N2 calibration 82 for an H ii region 200 pc away from SN Refsdal s site. From the same observations, the authors find a combined upper limit of 12 + log(o/h) < 8.11 dex from observations of nine H ii regions at nuclear offsets between 5 and 7 kpc 83, which is similar to the offsets of LS1 and SN Refsdal. 56

57 Recent work 84 has analyzed WFC3 grism spectra taken by GLASS 47,48 and follow-up observations of SN Refsdal. They fit the abundance measurement using a linear model, 12 + log(o/h) = ( ± ) r ± dex, (3) where r is the offset from the nucleus in kpc. This yields an abundance at LS1 s offset (assuming 7.9 ± 0.5 kpc) of 12 + log(o/h) = 8.29 ± 0.19 dex. This analysis does not take into account the [N ii] line when estimating the oxygen abundance, as it can be a biased tracer at z > Finally, while [N ii] was not detected in the OSIRIS IFU spectra of the site of SN Refsdal 83, a 1 hr Keck-II MOSFIRE integration yielded a [N ii] detection. The [N ii] line strength yields a PP04 N2 oxygen abundance of 12 + log(o/h) = 8.3 ± 0.1 dex 49, which is in agreement with the above estimate made using the Maiolino calibration 86 from WFC3 grism spectra. Given the above grism as well as MOSFIRE [N ii] metallicity estimates, we use an oxygen abundance of 12 + log(o/h) 8.3 dex as the metallicity of the massive stellar population near LS1 s coordinates. The Castelli & Kurucz 2004 stellar atmosphere models 17 are parameterised based on the Grevesse & Sauval solar oxygen abundance of 12 + log(o/h) = 8.83±0.06 dex. Therefore, we adopt log(z/z ) = 0.5 when drawing comparisons with the Castelli & Kurucz ATLAS9 models. of Ref. 88, K-correction and Distance Modulus. We calculate K-corrections following Equation 2 K = 2.5 log 10 (1 + z) + m AB F125W,syn mvega V,syn, (4) 57

58 where z = 1.49, m AB F125W,syn is the WFC3 F125W synthetic magnitude of a redshifted model spectrum, and m Vega V,syn is the synthetic Johnson V-band magnitude of the rest-frame model spectrum. Here the K-correction K xy is defined as m y = M x + dm + K xy, (5) where m y is the observer-frame apparent magnitude in the y band, M x is the rest-frame absolute magnitude in the x band, and dm is the distance modulus. Using the best-fitting spectral models, we calculate K V,F125W = 1.10 mag, and adopt dm = mag at z = 1.49 (with no correction for magnification). Stellar-Mass Density Along the Line of Sight to LS1. We computed two separate estimates of the stellar-mass density to LS1. The first estimate was the value we used when we created most of the simulated light curves, but it excluded light from the nearby brightest cluster galaxy (BCG). We computed a second, improved estimate that accounted for all intracluster light (ICL) along the line of sight. The updated analysis yielded a density approximately twice as high as the initial value. Initial Estimate: Galaxies with m F160W < 26 AB mag are selected and fit with single Sérsic profiles by using GALFIT 89 in a postage stamp (300 pix 300 pix). At the same time, the local sky background, assumed to be constant across the stamp, is fitted with the galaxy light profile. After fitting all the galaxies in the field, we reconstruct the ICL map by using the estimated local sky background values. Overlapping pixels are stacked, weighted by the χ 2 ν value from the fit. 58

59 The uncertainty is estimated from the original root-mean-square (RMS) map (published by the Hubble Frontier Fields team) and the systematic differences caused by changing the stamp size. We repeat this procedure for the ACS WFC F435W, F606W, F814W, and WFC3 IR F105W, F125W, F140W, and F160W filter bands 90. A correction is applied for Galactic extinction 91. Stellar mass is estimated in each pixel using the Fitting and Assessment of Synthetic Templates (FAST) software tool 92 with the BC03 93 stellar population model. FAST uses the Galaxy Spectral Evolution Library (GALAXEV; code to assemble composite stellar populations. We use the BC03 isochrones, a Chabrier IMF, and an exponentially declining star-formation history. Stars have initial masses that are between 0.1 and 100 M, and models are computed for metallicities of 0.004, 0.008, 0.02, and The hot gas in the galaxy-cluster ICM is thought to destroy dust, and extinction from dust is assumed to be zero. We note, however, that allowing the dust extinction to be a free parameter would change the estimated stellar mass by < 1%. Revised Estimate: We calculated a second estimate for the stellar-mass density that includes the contribution of stellar light associated with the BCG. We first constructed a total of eight apertures around the BCG shown in Extended Data Fig. 12. The apertures offsets from the BCG center and F140W surface brightnesses are similar to those of LS1, and were selected to exclude point sources and cluster-member galaxies (except the BCG). We estimate ACS WFC F435W, F606W, F814W, and WFC3 IR F105W, F125W, F140W, and F160W fluxes within each aperture, and apply a correction for Galactic extinction 91. We next 59

60 determine the ratio between the stellar mass (M ) and the WFC3-IR F140W flux (L) within each aperture. We estimate M with FAST 92 and the BC03 93 stellar population synthesis models. We adopt a delayed exponentially declining star-formation history and include both subsolar and solar metallicity ( 0.02 and 0.008) populations. Separate model fits are made for Chabrier and Salpeter IMFs, and the stars in our BC03 population synthesis models have initial masses that are between 0.1 and 100 M. Within each aperture, the statistical uncertainties of the total WFC3-IR flux in each bandpass are 0.5%. Among fits within the apertures, the average e-folding time is 600 Myr, and, at redshift z = 0.54, the stellar population ages are, on average, 4 Gyr. The uncertainty in M /L for each aperture is 30%, which approximately equals the standard deviation among the best-fitting estimates for all apertures. To estimate the stellar-mass density along the line of sight to LS1, we multiply the mean M /L computed across all eight apertures by the average F140W surface brightness in the two apertures adjacent to LS1. These apertures adjacent to LS1 may contain contamination in observerframe optical bandpasses from the underlying, young lensed galaxy. Within the WFC3-IR F140W bandpass, however, light from the cluster dominates. For Chabrier and Salpeter IMFs, the stellar mass densities computed using the BC03 model are M kpc 2 and M kpc 2, respectively. These revised estimates as well as our initial estimate for the density include remnants, whose masses are computed using the Renzini initial final mass function 63. The total local projected mass density inferred from 60

61 cluster models 7,8,27 is M kpc 2. 61

62 Rest Wavelength (Å) Region Adjacent to LS1 / Lev 2016A Age=7.2 Myr AV =0.3 mag χ 2 = " LS1 Flux (µjy) Observed Wavelength (Å) Magnitude (AB) Age [Myr] = AV [mag] = AV [mag] Age [Myr] AV [mag] Extended Data Fig. 1: Constraints on the age and dust extinction of the stellar population along the lensed arc adjacent to LS1. LS1 s flux was measured in the circular cyan aperture and the background measured inside of the red dashed aperture shown in upper-left panel. Upper-right panel plots the SED of the underlying arc measured inside the aperture outlined by a blue boundary, after subtracting the background measured in the aperture outlined in magenta. Bottom panel shows the posterior probability distributions of the age and extinction A V of the stellar population. Spectra of the host galaxy favor a gas-phase metallicity of Z 0.3. At such a metallicity, we find a bimodel posterior probability distribution with peaks at 8 and 35 Myr. An age of 8 Myr would be consistent with the age of a blue supergiant star. The stellar population synthesis model is constructed using the Padova isochrones 36,37, and we apply a Cardelli extinction law with R V =

63 1 µas 5 µas µ Extended Data Fig. 2: Distinct magnification patterns for respective counterimages Lev16B (upper panel) and LS1/Lev16A (lower panel) of LS1 within the source-plane host galaxy at redshift z = 1.49 from a ray-tracing simulation. Extensive regions of low magnification ( 100) for negativeparity image Lev16A could explain why it is undetected in HST imaging acquired in all except a single epoch acquired from 2004 through The map for positive-parity image LS1/Lev16A lacks such regions of extensive low magnification, and it always detected in deep imaging. Plotted angular scale is in the source plane, and one µarcsec in each panel corresponds to a physical pc at redshift z = If LS1 has an apparent transverse velocity of 1000 km s 1, it would travel 1 µarcsecond in 8.6 observer-frame years. These ray-tracing simulations are realistic if Lev16B and LS1/Lev16A are mutual counterimages offset by 0.13 on opposite sides of the galaxy cluster s critical curve in the image plane, and each of the counterimages has an average magnification of 600. The galaxy-cluster caustic, which is offset by 2.1 pc from these maps, is oriented parallel to the horizontal axes of each panel. The different patterns of magnification correspond to the parity of the image; Lev16B has negative parity, while LS1/Lev16A has positive parity. Here we have created a random realization of foreground intracluster stars and remnants having a mass-density ( M kpc 2 ) matching that we infer for a Salpeter IMF. 63

64 1ʺ LS1 Extended Data Fig. 3: Predicted position of LS1 in separate, full image of its host galaxy created by MACS J1149 galaxy-cluster lens. The cluster lens create three images of the host galaxy at redshift z = We detected LS1 adjacent to the critical curve separating two partial, merging images which have opposite parity. The third, full image shown here is at a greater distance from the cluster center, and it shows that LS1 lies close to the tip of a spiral arm. 64

65 Fraction of Peaks from Persistently Visible Star Nearby Galaxies ( = 2.53 ± 0.08; Bresolin+98) Stellar Luminosity Function Power-Law Index Extended Data Fig. 4: When a lensed star has a bright average apparent magnitude (F125W < 27.7 mag), it will also be responsible for almost all bright microlensing peaks (F125W < 26 mag) observed near the critical curve for simple assumptions. Here we plot the fraction of bright peaks caused by the bright star against the index of the stellar luminosity function. Since 99% of events likely arise from the luminous star, it is likely that Lev16B corresponds to the same star as LS1. However, this simulation randomly assigns positions to massive stars. To determine whether the observed clustering of massive stars could yield a greater probability that LS 1 / Lev16A and Lev16B are different stars, we carry out a simulation instead using the observed absolute magnitudes and positions of stars in the 30 Doradus cluster in the LMC from the SIMBAD catalog, and find a similarly low probability. 65

66 10 1 Probability One <26mag Event (50 Visits) Two <26mag Events (50 Visits) Persistent Star Image <27.7 mag Near Caustic <0.06 pc <27.7 mag Nearby Galaxies (Bresolin+98) 30 Doradus (Approximation) Luminosity Function Power-Law Index Extended Data Fig. 5: Dependence of the probability of observing highly magnified stellar images on the stellar luminosity function of the underlying arc. Panel shows probabilities of (a) bright microlensing events (F125W <26 mag; solid green and dotted pink), (b) a persistently bright magnified star (F125W <27.7 mag) similar to that observed at LS1 s position in (dashed brown), and (c) a persistently bright magnified star (F125W <27.7 mag) within 0.06 pc (dot-dash purple). Probabilities are small given the index of stellar luminosity function measured for nearby galaxies (α = 2.53 ± 0.08; vertical blue) 20, but become significantly larger for shallower powerlaw indices, such as that for the 30 Doradus star-forming region in the LMC (vertical orange; approximate). Here we have assumed N obs = 50 visits by HST, the number of separate observations of MACS J1149 taken through 13 April 2017 after binning data by 10 days. The lower stellar luminosity limit used for these simulations is 10 L. 66

67 Extended Data Fig. 6: Offsets from critical curve and luminosities of lensed stars for different luminosity functions from 10 6 Monte Carlo simulations. We use the surface brightness (F125W 25 mag arcsec 2 ) measured along the 0.2 -wide arc to constrain the normalization of the stellar luminosity function, and then Poisson statistics to populate the source plane. The lower luminosity limit used for these simulations is 10 L. Upper panels show that stars with F125W 27.7 magover a period lasting many years should only appear within 0.15 of the critical curve, and have luminosities of L. Lower panels show expected offset distribution of bright microlensing peaks (F125W 26 mag) to 0.4, and of the luminosities of lensed stars. A stellar luminosity function similar to that measured in nearby galaxies (α = 2.53 ± 0.08) yields fewer events with less-luminous stars 20. Left panels indicate the offset θ in arcseconds of stars from the critical curve, while right panels show the intrinsic luminosity of the stars in units of the solar luminosity (L ). 67

68 Flux (µjy) Lev 2016B Spera15 (χ 2 =112.6) Spera15/3% PBH (χ 2 =174.3) Magnitude (AB) Date (MJD) Flux (µjy) Lev 2017A Combined F125W Light Curve 27 Magnitude (AB) Date (MJD) Extended Data Fig. 7: Light curve at the positions of Lev16B (upper panel) detected on 30 October 2016 and potential event Lev 2017A (lower panel) detected on 3 January Fluxes measured through all wide-band HST filters are converted to F125W using LS1 s SED. Lev 2017A is only offset from Lev16B by 0.10, so flux measurements at their positions are correlated. The first (higher) peak in Lev 2017A s light curve plotted here corresponds to flux from Lev16B. 68

69 Magnification Magnification Magnification a b c Woosley Observer-Frame Years Fryer Observer-Frame Years Spera Observer-Frame Years Extended Data Fig. 8: Simulated light curves of a star at θ = 0.13 from cluster critical curve for three stellar evolution and core-collapse models. The mass functions are constructed using a Chabrier IMF and a prescription for binary fractions and mass ratios at low redshift 22. The stellar evolution models used to determine the initial final mass function for each star include the solar-metallicity, single stellar evolution models (Woosley02) 23, as well as single stars at subsolar metallicity (Z = 0.3 Z ; Z = 0.006) where BHs with masses up to 30 M form from the collapse of massive stars (Fryer12) 24, and stars having initial masses greater than 33 M become BHs with masses within the range M (Spera15) 25. The Fryer12 and Spera15 models contain greater numbers of BH remnants, which may yield a higher frequency of decade-long intervals with low magnification. 69

70 Magnification a Spera15 Chabrier (High-Mass Density) Magnification b Observer-Frame Years Spera15 Salpeter (High-Mass Density) Observer-Frame Years Extended Data Fig. 9: Simulated light curves of a star at θ = 0.13 from cluster critical curve for Chabrier (upper panel) and Salpeter (lower panel) IMFs. The simulated light curve constructed from the model with a Salpeter IMF yields a greater number of peaks. Plotted models are constructed using a prescription for binary fractions and mass ratios at low redshift 22. The stellarmass densities are M kpc 2 and M kpc 2 for the upper and lower plots, respectively, and are the best-fitting values to the SED of the ICL for stellar-population synthesis models constructed using Chabrier and Salpeter IMFs. 70

71 Magnification Magnification Magnification a b c Spera15 + 1% PBH Observer-Frame Years Spera15 + 3% PBH Observer-Frame Years Spera % PBH Observer-Frame Years Extended Data Fig. 10: Effect of increasing abundance of 30 M BHs on the simulated light curves of star at θ = Replacing 1% (upper panel), 3% (middle), and 10% (bottom) of smooth DM with 30 M BHs yields light curves where the average magnification varies on an increasingly long timescales. An extended period of low magnification for one of the pair of images could help to explain why only a single image, LS1 / Lev16A, is persistently visible in HST imaging. 71

72 Relative Number Density Mass (M ) All Stars Single Chabrier IMF Woosley02 Fryer12 Spera15 Relative Number Density Binary Prescription Chabrier IMF Woosley02 Fryer12 Spera Mass (M ) Extended Data Fig. 11: Differences among the mass distributions of surviving stars and stellar remnants (i.e., white dwarf stars, neutron stars, and BHs) for the Woosley02, Fryer12, and Spera15 stellar evolution models. Left panel plots the mass distributions assuming no stars have companions, and right panel shows mass functions assuming the mass-dependent binary fractions and mass ratios measured in the nearby universe

73 3" LS1 Extended Data Fig. 12: Apertures used for revised estimate of galaxy-cluster stellar-mass density along line-of-sight to LS1. We first estimate the mean M /L across all eight apertures using stellarpopulation synthesis models. We next multiply the average F140W surface brightness in the two apertures adjacent to LS1 by the mean value of M /L to estimate the stellar mass density along the line of sight to LS1. 73

74 100 2 (Generative - Best-Fit Models) Average 2 of Model Fits Extended Data Fig. 13: Confidence intervals for χ 2 statistics determined using simulated light curves. For the models listed in Table 1, we generate simulated light curves for stars with M V = { 8, 9, 10}, and fit them, allowing the lensed star to have an absolute magnitude within the range 7.5 < M V < 9.5. The dashed black vertical shows the average of the χ 2 statistics for the Table 1 models for LS 1/ Lev16A and Lev16B. For all simulated light curves where the average χ 2 value is within 100 of the vertical dashed line, we calculate the difference χ 2 values between the χ 2 values of the generative ( true ) model and of the best-fitting model. For 68% of simulated light curves, χ 2 13, and, for 95% of simulated light curves, χ

75 Extended Data Fig. 14: Example of trains of multiple counterimages of a single background star as it traverses the region near the galaxy-cluster caustic. In general, there is a single counterimage per microlens 94, but most images have magnifications of order unity or even less and are not detectable. We show example of trains without PBHs (left panel), where 30 M PBHs account for 1% of DM (middle panel), and where 30 M PBHs account for 10% of DM (right panel). Increasing the PBH abundance yields more extended trains, although their extent ( 3 milliarcsec) when PBHs account for 10% would be too small to detect in HST imaging. The simulation shown is of a 1000 R star whose image appears at an offset of 0.13 from the cluster critical curve. Near peak magnification, the star appears as a train of counterimages. Each cirlce in right panel encloses one image, and each ellipse encloses a set of two or three closely spaced images, in the train. The sizes of the circles and ellipses indicate the magnification of each image or set of images, respectively. In addition to replacing fractions of cluster DM with PBHs, we have populated the lens plane with stars and compact object remnants to match the mass density in surviving stars and remnants we infer for the ICM ( M kpc 2 ). 75

Ultra high magnification microlensing

Ultra high magnification microlensing Ultra high magnification microlensing Masamune Oguri (University of Tokyo) 2018/2/16 PACIFIC 2018@Kiroro Dark Matter (DM) 1/4 of the mass of the Universe is composed of dark matter a lot of evidence from

More information

Source plane reconstruction of the giant gravitational arc in Abell 2667: a condidate Wolf-Rayet galaxy at z 1

Source plane reconstruction of the giant gravitational arc in Abell 2667: a condidate Wolf-Rayet galaxy at z 1 Source plane reconstruction of the giant gravitational arc in Abell 2667: a condidate Wolf-Rayet galaxy at z 1 Speaker: Shuo Cao Department of Astronomy Beijing Normal University Collaborators: Giovanni

More information

Multiple Images of a Highly Magnified Supernova Formed by an Early-Type Cluster Galaxy Lens

Multiple Images of a Highly Magnified Supernova Formed by an Early-Type Cluster Galaxy Lens arxiv:1411.6009v2 [astro-ph.co] 24 Nov 2014 Multiple Images of a Highly Magnified Supernova Formed by an Early-Type Cluster Galaxy Lens Patrick L. Kelly 1, Steven A. Rodney 2, Tommaso Treu 3, Ryan J. Foley

More information

Exploring the Depths of the Universe

Exploring the Depths of the Universe Exploring the Depths of the Universe Jennifer Lotz Hubble Science Briefing Jan. 16, 2014 Hubble is now observing galaxies 97% of the way back to the Big Bang, during the first 500 million years 2 Challenge:

More information

Supplementary Information for SNLS-03D3bb a super- Chandrasekhar mass Type Ia supernova

Supplementary Information for SNLS-03D3bb a super- Chandrasekhar mass Type Ia supernova 1 Supplementary Information for SNLS-03D3bb a super- Chandrasekhar mass Type Ia supernova SN Location SNLS-03D3bb is located at RA: 14:16:18.920 Dec: +52:14:53.66 (J2000) in the D3 (extended Groth Strip)

More information

Type II Supernovae as Standardized Candles

Type II Supernovae as Standardized Candles Type II Supernovae as Standardized Candles Mario Hamuy 1 2 Steward Observatory, The University of Arizona, Tucson, AZ 85721 Philip A. Pinto Steward Observatory, The University of Arizona, Tucson, AZ 85721

More information

Searching primeval galaxies through gravitational telescopes

Searching primeval galaxies through gravitational telescopes Mem. S.A.It. Suppl. Vol. 19, 258 c SAIt 2012 Memorie della Supplementi Searching primeval galaxies through gravitational telescopes A. Monna 1 and G. Covone 1,2,3 1 Dipartimento di Scienze Fisiche, Università

More information

A Look Back: Galaxies at Cosmic Dawn Revealed in the First Year of the Hubble Frontier Fields Initiative

A Look Back: Galaxies at Cosmic Dawn Revealed in the First Year of the Hubble Frontier Fields Initiative A Look Back: Galaxies at Cosmic Dawn Revealed in the First Year of the Hubble Frontier Fields Initiative Dr. Gabriel Brammer (ESA/AURA, STScI) Hubble Science Briefing / November 6, 2014 1 The Early Universe

More information

3/6/12! Astro 358/Spring 2012! Galaxies and the Universe! Dark Matter in Spiral Galaxies. Dark Matter in Galaxies!

3/6/12! Astro 358/Spring 2012! Galaxies and the Universe! Dark Matter in Spiral Galaxies. Dark Matter in Galaxies! 3/6/12 Astro 358/Spring 2012 Galaxies and the Universe Dark Matter in Galaxies Figures + Tables for Lectures (Feb 16-Mar 6) Dark Matter in Spiral Galaxies Flat rotation curve of Milky Way at large radii

More information

arxiv: v1 [astro-ph.ga] 27 Oct 2015

arxiv: v1 [astro-ph.ga] 27 Oct 2015 Luminosity functions in the CLASH-VLT cluster MACS J1206.2-0847: the importance of tidal interactions arxiv:1510.08097v1 [astro-ph.ga] 27 Oct 2015 A. Mercurio 1, M. Annunziatella 2,3, A. Biviano 3, M.

More information

The SINFONI Nearby Elliptical Lens Locator Survey (SNELLS)

The SINFONI Nearby Elliptical Lens Locator Survey (SNELLS) The SINFONI Nearby Elliptical Lens Locator Survey (SNELLS) Russell J. Smith 1 John R. Lucey 1 Charlie Conroy 2 1 Centre for Extragalactic Astronomy, Durham University, United Kingdom 2 Harvard Smithsonian

More information

Resolved Star Formation Surface Density and Stellar Mass Density of Galaxies in the Local Universe. Abstract

Resolved Star Formation Surface Density and Stellar Mass Density of Galaxies in the Local Universe. Abstract Resolved Star Formation Surface Density and Stellar Mass Density of Galaxies in the Local Universe Abdurrouf Astronomical Institute of Tohoku University Abstract In order to understand how the stellar

More information

BUILDING GALAXIES. Question 1: When and where did the stars form?

BUILDING GALAXIES. Question 1: When and where did the stars form? BUILDING GALAXIES The unprecedented accuracy of recent observations of the power spectrum of the cosmic microwave background leaves little doubt that the universe formed in a hot big bang, later cooling

More information

Stellar Populations: Resolved vs. unresolved

Stellar Populations: Resolved vs. unresolved Outline Stellar Populations: Resolved vs. unresolved Individual stars can be analyzed Applicable for Milky Way star clusters and the most nearby galaxies Integrated spectroscopy / photometry only The most

More information

Number of Stars: 100 billion (10 11 ) Mass : 5 x Solar masses. Size of Disk: 100,000 Light Years (30 kpc)

Number of Stars: 100 billion (10 11 ) Mass : 5 x Solar masses. Size of Disk: 100,000 Light Years (30 kpc) THE MILKY WAY GALAXY Type: Spiral galaxy composed of a highly flattened disk and a central elliptical bulge. The disk is about 100,000 light years (30kpc) in diameter. The term spiral arises from the external

More information

PoS(extremesky2009)103

PoS(extremesky2009)103 The study of the nature of sources AX J1749.1 2733 and AX J1749.2 2725 D. I. Karasev Space Research Institute, Profsoyuznaya str. 84/32, Moscow 117997, Russia E-mail: dkarasev@iki.rssi.ru Space Research

More information

DARK MATTER UNDER THE MICROSCOPE: CONSTRAINING COMPACT DARK MATTER WITH CAUSTIC CROSSING EVENTS

DARK MATTER UNDER THE MICROSCOPE: CONSTRAINING COMPACT DARK MATTER WITH CAUSTIC CROSSING EVENTS Draft version February 22, 2018 Preprint typeset using L A TEX style emulateapj v. 12/16/11 DARK MATTER UNDER THE MICROSCOPE: CONSTRAINING COMPACT DARK MATTER WITH CAUSTIC CROSSING EVENTS Jose M. Diego

More information

arxiv: v2 [astro-ph.co] 23 Mar 2018

arxiv: v2 [astro-ph.co] 23 Mar 2018 Draft version March 28, 2018 Preprint typeset using L A TEX style emulateapj v. 12/16/11 DARK MATTER UNDER THE MICROSCOPE: CONSTRAINING COMPACT DARK MATTER WITH CAUSTIC CROSSING EVENTS Jose M. Diego 1,

More information

High Redshift Universe

High Redshift Universe High Redshift Universe Finding high z galaxies Lyman break galaxies (LBGs) Photometric redshifts Deep fields Starburst galaxies Extremely red objects (EROs) Sub-mm galaxies Lyman α systems Finding high

More information

The sizes of z ~ 6-8 lensed galaxies from the Hubble Frontier Fields data of Abell 2744

The sizes of z ~ 6-8 lensed galaxies from the Hubble Frontier Fields data of Abell 2744 The sizes of z ~ 6-8 lensed galaxies from the Hubble Frontier Fields data of Abell 2744 Kawamata+15, ApJ, 804, 103 Ryota Kawamata The University of Tokyo With: Masafumi Ishigaki, Kazuhiro Shimasaku, Masamune

More information

Learning Objectives: Chapter 13, Part 1: Lower Main Sequence Stars. AST 2010: Chapter 13. AST 2010 Descriptive Astronomy

Learning Objectives: Chapter 13, Part 1: Lower Main Sequence Stars. AST 2010: Chapter 13. AST 2010 Descriptive Astronomy Chapter 13, Part 1: Lower Main Sequence Stars Define red dwarf, and describe the internal dynamics and later evolution of these low-mass stars. Appreciate the time scale of late-stage stellar evolution

More information

COSMOLOGY PHYS 30392 OBSERVING THE UNIVERSE Part I Giampaolo Pisano - Jodrell Bank Centre for Astrophysics The University of Manchester - January 2013 http://www.jb.man.ac.uk/~gp/ giampaolo.pisano@manchester.ac.uk

More information

Beyond Our Solar System Chapter 24

Beyond Our Solar System Chapter 24 Beyond Our Solar System Chapter 24 PROPERTIES OF STARS Distance Measuring a star's distance can be very difficult Stellar parallax Used for measuring distance to a star Apparent shift in a star's position

More information

Hubble s Law and the Cosmic Distance Scale

Hubble s Law and the Cosmic Distance Scale Lab 7 Hubble s Law and the Cosmic Distance Scale 7.1 Overview Exercise seven is our first extragalactic exercise, highlighting the immense scale of the Universe. It addresses the challenge of determining

More information

Part two of a year-long introduction to astrophysics:

Part two of a year-long introduction to astrophysics: ASTR 3830 Astrophysics 2 - Galactic and Extragalactic Phil Armitage office: JILA tower A909 email: pja@jilau1.colorado.edu Spitzer Space telescope image of M81 Part two of a year-long introduction to astrophysics:

More information

Age-redshift relation. The time since the big bang depends on the cosmological parameters.

Age-redshift relation. The time since the big bang depends on the cosmological parameters. Age-redshift relation The time since the big bang depends on the cosmological parameters. Lyman Break Galaxies High redshift galaxies are red or absent in blue filters because of attenuation from the neutral

More information

Exploiting Cosmic Telescopes with RAVEN

Exploiting Cosmic Telescopes with RAVEN Exploiting Cosmic Telescopes with RAVEN S. Mark Ammons Lawrence Livermore National Laboratory Thanks to: Ken Wong (Arizona) Ann Zabludoff (Arizona) Chuck Keeton (Rutgers) K. Decker French (Arizona) RAVEN

More information

Windows to the Past: Using Gravitational Telescopes to Study our Cosmic Origins

Windows to the Past: Using Gravitational Telescopes to Study our Cosmic Origins Windows to the Past: Using Gravitational Telescopes to Study our Cosmic Origins Austin Hoag University of California, Davis July 15, 2016 1 / 21 Collaborators Maru sa Brada c (PhD advisor; UCD), Tommaso

More information

Supernova Explosions. Novae

Supernova Explosions. Novae Supernova Explosions Novae Novae occur in close binary-star systems in which one member is a white dwarf. First, mass is transferred from the normal star to the surface of its white dwarf companion. 1

More information

The First Galaxies: Evolution drivers via luminosity functions and spectroscopy through a magnifying GLASS

The First Galaxies: Evolution drivers via luminosity functions and spectroscopy through a magnifying GLASS Charlotte Mason (UCLA) Aspen, 7 Feb 2016 The First Galaxies: Evolution drivers via luminosity functions and spectroscopy through a magnifying GLASS with Tommaso Treu (UCLA), Michele Trenti (U. Melbourne),

More information

Quasars and Active Galactic Nuclei (AGN)

Quasars and Active Galactic Nuclei (AGN) Quasars and Active Galactic Nuclei (AGN) Astronomy Summer School in Mongolia National University of Mongolia, Ulaanbaatar July 21-26, 2008 Kaz Sekiguchi Hubble Classification M94-Sa M81-Sb M101-Sc M87-E0

More information

Cooking with Strong Lenses and Other Ingredients

Cooking with Strong Lenses and Other Ingredients Cooking with Strong Lenses and Other Ingredients Adam S. Bolton Department of Physics and Astronomy The University of Utah AASTCS 1: Probes of Dark Matter on Galaxy Scales Monterey, CA, USA 2013-July-17

More information

Radio Nebulae around Luminous Blue Variable Stars

Radio Nebulae around Luminous Blue Variable Stars Radio Nebulae around Luminous Blue Variable Stars Claudia Agliozzo 1 G. Umana 2 C. Trigilio 2 C. Buemi 2 P. Leto 2 A. Ingallinera 1 A. Noriega-Crespo 3 J. Hora 4 1 University of Catania, Italy 2 INAF-Astrophysical

More information

PESSTO The Public ESO Spectroscopic Survey for Transient Objects

PESSTO The Public ESO Spectroscopic Survey for Transient Objects PESSTO The Public ESO Spectroscopic Survey for Transient Objects Morgan Fraser Institute of Astronomy, University of Cambridge, UK Stefano Benetti INAF - Osservatorio Astronomico di Padova, Italy Cosimo

More information

Unravelling the progenitors of merging black hole binaries

Unravelling the progenitors of merging black hole binaries Unravelling the progenitors of merging black hole binaries Dipartimento di Fisica e Astronomia G. Galilei, Università di Padova. INAF, Osservatorio Astronomico di Padova, Padova, Italy. INFN, Milano Bicocca,

More information

Cosmology at a Crossroads: Tension With the Hubble Constant

Cosmology at a Crossroads: Tension With the Hubble Constant Cosmology at a Crossroads: Tension With the Hubble Constant Wendy L. Freedman We are at an interesting juncture in cosmology. With new methods and technology, the accuracy in measurement of the Hubble

More information

Dark Matter ASTR 2120 Sarazin. Bullet Cluster of Galaxies - Dark Matter Lab

Dark Matter ASTR 2120 Sarazin. Bullet Cluster of Galaxies - Dark Matter Lab Dark Matter ASTR 2120 Sarazin Bullet Cluster of Galaxies - Dark Matter Lab Mergers: Test of Dark Matter vs. Modified Gravity Gas behind DM Galaxies DM = location of gravity Gas = location of most baryons

More information

2019 Astronomy Team Selection Test

2019 Astronomy Team Selection Test 2019 Astronomy Team Selection Test Acton-Boxborough Regional High School Written by Antonio Frigo Do not flip over this page until instructed. Instructions You will have 45 minutes to complete this exam.

More information

3D spectroscopy of massive stars, SNe, and other point sources in crowded fields

3D spectroscopy of massive stars, SNe, and other point sources in crowded fields 3D spectroscopy of massive stars, SNe, and other point sources in crowded fields Martin M. Roth Sebastian Kamann, Christer Sandin, Ana Monreal, Peter Weilbacher, Lutz Wisotzki Leibniz-Institut für Astrophysik

More information

Survey of Astrophysics A110

Survey of Astrophysics A110 Goals: Galaxies To determine the types and distributions of galaxies? How do we measure the mass of galaxies and what comprises this mass? How do we measure distances to galaxies and what does this tell

More information

1. The AGB dust budget in nearby galaxies

1. The AGB dust budget in nearby galaxies **Volume Title** ASP Conference Series, Vol. **Volume Number** **Author** c **Copyright Year** Astronomical Society of the Pacific Identifying the chemistry of the dust around AGB stars in nearby galaxies

More information

arxiv:astro-ph/ v1 13 Apr 2006

arxiv:astro-ph/ v1 13 Apr 2006 **FULL TITLE** ASP Conference Series, Vol. **VOLUME**, **YEAR OF PUBLICATION** **NAMES OF EDITORS** Clusters of Galaxies at 1 < z < 2 : The Spitzer Adaptation of the Red-Sequence Cluster Survey arxiv:astro-ph/0604289

More information

PoS(LCDU 2013)010. The extinction law at high redshift. Simona Gallerani Scuola Normale Superiore

PoS(LCDU 2013)010. The extinction law at high redshift. Simona Gallerani Scuola Normale Superiore Scuola Normale Superiore E-mail: simona.gallerani@sns.it We analyze the optical-near infrared spectra of 33 quasars with redshifts 3.9 < z < 6.4 with the aim of investigating the properties of dust extinction

More information

SN-MCT Science Goals

SN-MCT Science Goals THE CANDELS+CLASH SUPERNOVA PROGRAM Adam Riess (STScI/JHU) Azalee Bostroem Stefano Casertano Brad Cenko Pete Challis Tomas Dahlen Mark Dickinson Harry Ferguson Alex Filippenko Peter Garnavich Or Graur

More information

Review of results from the EROS microlensing search for massive compact objects

Review of results from the EROS microlensing search for massive compact objects Author manuscript, published in "IDM008 - identification of dark matter 008, Stockholm : Sweden (008)" Review of results from the EROS microlensing search for massive compact objects Laboratoire de l Accélérateur

More information

Chapter 14 The Milky Way Galaxy

Chapter 14 The Milky Way Galaxy Chapter 14 The Milky Way Galaxy Spiral Galaxy M81 - similar to our Milky Way Galaxy Our Parent Galaxy A galaxy is a giant collection of stellar and interstellar matter held together by gravity Billions

More information

Searching for black holes in nearby galaxies with Simbol-X

Searching for black holes in nearby galaxies with Simbol-X Mem. S.A.It. Vol. 79, 208 c SAIt 2008 Memorie della Searching for black holes in nearby galaxies with Simbol-X Paul Gorenstein Harvard-Smithsonian Center for Astrophysics, 60 Garden St. Cambridge, MA 02138,

More information

Astro2010 Science White Paper: Tracing the Mass Buildup of Supermassive Black Holes and their Host Galaxies

Astro2010 Science White Paper: Tracing the Mass Buildup of Supermassive Black Holes and their Host Galaxies Astro2010 Science White Paper: Tracing the Mass Buildup of Supermassive Black Holes and their Host Galaxies Anton M. Koekemoer (STScI) Dan Batcheldor (RIT) Marc Postman (STScI) Rachel Somerville (STScI)

More information

New perspectives on red supergiants

New perspectives on red supergiants Highlights on Spanish Astrophysics IX, Proceedings of the XII Scientific Meeting of the Spanish Astronomical Society held on July 1 22, 201, in Bilbao, Spain. S. Arribas, A. Alonso-Herrero, F. Figueras,

More information

9. Evolution with redshift - z > 1.5. Selection in the rest-frame UV

9. Evolution with redshift - z > 1.5. Selection in the rest-frame UV 11-5-10see http://www.strw.leidenuniv.nl/ franx/college/galaxies10 10-c09-1 11-5-10see http://www.strw.leidenuniv.nl/ franx/college/galaxies10 10-c09-2 9. Evolution with redshift - z > 1.5 Selection in

More information

Gamma-Ray Astronomy. Astro 129: Chapter 1a

Gamma-Ray Astronomy. Astro 129: Chapter 1a Gamma-Ray Bursts Gamma-Ray Astronomy Gamma rays are photons with energies > 100 kev and are produced by sub-atomic particle interactions. They are absorbed by our atmosphere making observations from satellites

More information

Chapter 10: Unresolved Stellar Populations

Chapter 10: Unresolved Stellar Populations Chapter 10: Unresolved Stellar Populations We now consider the case when individual stars are not resolved. So we need to use photometric and spectroscopic observations of integrated magnitudes, colors

More information

Supernova Explosions. Novae

Supernova Explosions. Novae Supernova Explosions Novae Novae occur in close binary-star systems in which one member is a white dwarf. First, mass is transferred from the normal star to the surface of its white dwarf companion. 1

More information

The phenomenon of gravitational lenses

The phenomenon of gravitational lenses The phenomenon of gravitational lenses The phenomenon of gravitational lenses If we look carefully at the image taken with the Hubble Space Telescope, of the Galaxy Cluster Abell 2218 in the constellation

More information

We investigate the flux ratio anomalies between

We investigate the flux ratio anomalies between arxiv:1711.07919v1 [astro-ph.co] 21 Nov 2017 A Quadruply Lensed SN Ia: Gaining a Time-Delay...Losing a Standard Candle Daniel A. Yahalomi 1, Paul L. Schechter 1,2, and Joachim Wambsganss 3 1 MIT Department

More information

Gravitational Lensing. A Brief History, Theory, and Applications

Gravitational Lensing. A Brief History, Theory, and Applications Gravitational Lensing A Brief History, Theory, and Applications A Brief History Einstein (1915): light deflection by point mass M due to bending of space-time = 2x Newtonian light tangentially grazing

More information

Age Dating A SSP. Quick quiz: please write down a 3 sentence explanation of why these plots look like they do.

Age Dating A SSP. Quick quiz: please write down a 3 sentence explanation of why these plots look like they do. Color is only a weak function of age after ~3Gyrs (for a given metallicity) (See MBW pg 473) But there is a strong change in M/L V and weak change in M/L K Age Dating A SSP Quick quiz: please write down

More information

Hubble Space Telescope Frontier Fields MidTerm Review

Hubble Space Telescope Frontier Fields MidTerm Review Hubble Space Telescope Frontier Fields MidTerm Review *Membership: James Bullock (UC-Irvine) [Chair], Mark Dickinson (NOAO), Richard Ellis (Caltech), Mariska Kriek (UC-Berkeley), Sally Oey (U. Michigan),

More information

SUPPLEMENTARY INFORMATION

SUPPLEMENTARY INFORMATION 1. Identification of classical Cepheids: We identified three classical Cepheids amongst the 45 short-period variables discovered. Our sample includes classical Cepheids, type II Cepheids, eclipsing binaries

More information

Spectroscopy of Blue Supergiants in the Disks of Spiral Galaxies: Metallicities and Distances. Rolf Kudritzki

Spectroscopy of Blue Supergiants in the Disks of Spiral Galaxies: Metallicities and Distances. Rolf Kudritzki Spectroscopy of Blue Supergiants in the Disks of Spiral Galaxies: Metallicities and Distances Rolf Kudritzki ΛCDM-universe metallicity of galaxies depends on their mass Extragalactic stellar astronomy

More information

GALACTIC Al 1.8 MeV GAMMA-RAY SURVEYS WITH INTEGRAL

GALACTIC Al 1.8 MeV GAMMA-RAY SURVEYS WITH INTEGRAL Proceedings of the 3rd Galileo Xu Guangqi Meeting International Journal of Modern Physics: Conference Series Vol. 23 (2013) 48 53 c World Scientific Publishing Company DOI: 10.1142/S2010194513011069 GALACTIC

More information

CAULDRON: dynamics meets gravitational lensing. Matteo Barnabè

CAULDRON: dynamics meets gravitational lensing. Matteo Barnabè CAULDRON: dynamics meets gravitational lensing Matteo Barnabè KIPAC/SLAC, Stanford University Collaborators: Léon Koopmans (Kapteyn), Oliver Czoske (Vienna), Tommaso Treu (UCSB), Aaron Dutton (MPIA), Matt

More information

3 The lives of galaxies

3 The lives of galaxies Discovering Astronomy : Galaxies and Cosmology 24 3 The lives of galaxies In this section, we look at how galaxies formed and evolved, and likewise how the large scale pattern of galaxies formed. But before

More information

Astro 101 Slide Set: Multiple Views of an Extremely Distant Galaxy

Astro 101 Slide Set: Multiple Views of an Extremely Distant Galaxy Astro 101 Slide Set: Multiple Views of an Extremely Distant Galaxy A Discovery from the Hubble Frontier Fields Program Topic: Distant galaxies Concepts: Galaxy development, Gravitational lensing Missions:

More information

The Red Supergiant Progenitors of Core-collapse Supernovae. Justyn R. Maund

The Red Supergiant Progenitors of Core-collapse Supernovae. Justyn R. Maund The Red Supergiant Progenitors of Core-collapse Supernovae Justyn R. Maund Stellar Populations Workshop IAG-USP 1 st December 2015 Astronomy and Astrophysics at Sheffield 8 Academic staff 5 Postdocs 13

More information

Testing the cold dark matter model with gravitational lensing

Testing the cold dark matter model with gravitational lensing Testing the cold dark matter model with gravitational lensing Priyamvada Natarajan Departments of Astronomy & Physics Yale University In Memoriam: Vera Rubin New Directions in Theoretical Physics, January

More information

Hubble Space Telescope ultraviolet spectroscopy of blazars: emission lines properties and black hole masses. E. Pian, R. Falomo, A.

Hubble Space Telescope ultraviolet spectroscopy of blazars: emission lines properties and black hole masses. E. Pian, R. Falomo, A. Hubble Space Telescope ultraviolet spectroscopy of blazars: emission lines properties and black hole masses E. Pian, R. Falomo, A. Treves 1 Outline Extra Background Introduction Sample Selection Data Analysis

More information

Physics of Galaxies 2016 Exercises with solutions batch I

Physics of Galaxies 2016 Exercises with solutions batch I Physics of Galaxies 2016 Exercises with solutions batch I 1. Distance and brightness at low redshift You discover an interesting galaxy in the local Universe and measure its redshift to be z 0.053 and

More information

EVOLUTION OF DUST EXTINCTION AND SUPERNOVA COSMOLOGY

EVOLUTION OF DUST EXTINCTION AND SUPERNOVA COSMOLOGY To Appear in ApJ Letters Preprint typeset using LATEX style emulateapj v. 04/03/99 EVOLUTION OF DUST EXTINCTION AND SUPERNOVA COSMOLOGY Tomonori Totani 1 and Chiaki Kobayashi 2 1 National Astronomical

More information

Our Galaxy. We are located in the disk of our galaxy and this is why the disk appears as a band of stars across the sky.

Our Galaxy. We are located in the disk of our galaxy and this is why the disk appears as a band of stars across the sky. Our Galaxy Our Galaxy We are located in the disk of our galaxy and this is why the disk appears as a band of stars across the sky. Early attempts to locate our solar system produced erroneous results.

More information

Dust properties of galaxies at redshift z 5-6

Dust properties of galaxies at redshift z 5-6 Dust properties of galaxies at redshift z 5-6 Ivana Barisic 1, Supervisor: Dr. Peter L. Capak 2, and Co-supervisor: Dr. Andreas Faisst 2 1 Physics Department, University of Zagreb, Zagreb, Croatia 2 Infrared

More information

arxiv:astro-ph/ v1 23 Dec 2005

arxiv:astro-ph/ v1 23 Dec 2005 3D spectroscopy as a tool for investigation of the BLR of lensed QSOs Luka Č. Popović Astronomical Observatory, Volgina 7, 11160 Belgrade, Serbia lpopovic@aob.bg.ac.yu arxiv:astro-ph/0512594v1 23 Dec 2005

More information

The cosmic distance scale

The cosmic distance scale The cosmic distance scale Distance information is often crucial to understand the physics of astrophysical objects. This requires knowing the basic properties of such an object, like its size, its environment,

More information

TEREZA JEŘÁBKOVÁ ESO GARCHING & UNIVERSITY OF BONN & CHARLES UNIVERSITY IN PRAGUE

TEREZA JEŘÁBKOVÁ ESO GARCHING & UNIVERSITY OF BONN & CHARLES UNIVERSITY IN PRAGUE 1 TEREZA JEŘÁBKOVÁ ESO GARCHING & UNIVERSITY OF BONN & CHARLES UNIVERSITY IN PRAGUE www: sirrah.troja.mff.cuni.cz/~tereza email: tjerabko@eso.org 17 MODEST UNDER PRAGUE S STARRY SKIES CZECH REPUBLIC 18-22

More information

Our Galaxy. Milky Way Galaxy = Sun + ~100 billion other stars + gas and dust. Held together by gravity! The Milky Way with the Naked Eye

Our Galaxy. Milky Way Galaxy = Sun + ~100 billion other stars + gas and dust. Held together by gravity! The Milky Way with the Naked Eye Our Galaxy Milky Way Galaxy = Sun + ~100 billion other stars + gas and dust Held together by gravity! The Milky Way with the Naked Eye We get a special view of our own galaxy because we are part of it!

More information

LECTURE 1: Introduction to Galaxies. The Milky Way on a clear night

LECTURE 1: Introduction to Galaxies. The Milky Way on a clear night LECTURE 1: Introduction to Galaxies The Milky Way on a clear night VISIBLE COMPONENTS OF THE MILKY WAY Our Sun is located 28,000 light years (8.58 kiloparsecs from the center of our Galaxy) in the Orion

More information

Supernovae explosions and the Accelerating Universe. Bodo Ziegler

Supernovae explosions and the Accelerating Universe. Bodo Ziegler Nobel Prize for Physics 2011 Supernovae explosions and the Accelerating Universe Institute for Astronomy University of Vienna Since 09/2010: ouniprof University of Vienna 12/2008-08/10: Staff member European

More information

arxiv: v1 [astro-ph.ga] 8 Jul 2017

arxiv: v1 [astro-ph.ga] 8 Jul 2017 Two Peculiar Fast Transients in a Strongly Lensed Host Galaxy arxiv:1707.02434v1 [astro-ph.ga] 8 Jul 2017 S. A. Rodney 1, I. Balestra 2, M. Bradač 3, G. Brammer 4, T. Broadhurst 5,6, G. B. Caminha 7, G.

More information

Dark Matter. Galaxy Counts Redshift Surveys Galaxy Rotation Curves Cluster Dynamics Gravitational Lenses ~ 0.3 Ω M Ω b.

Dark Matter. Galaxy Counts Redshift Surveys Galaxy Rotation Curves Cluster Dynamics Gravitational Lenses ~ 0.3 Ω M Ω b. Dark Matter Galaxy Counts Redshift Surveys Galaxy Rotation Curves Cluster Dynamics Gravitational Lenses Ω M ~ 0.3 2 1 Ω b 0.04 3 Mass Density by Direct Counting Add up the mass of all the galaxies per

More information

THE GALACTIC BULGE AS SEEN BY GAIA

THE GALACTIC BULGE AS SEEN BY GAIA 143 THE GALACTIC BULGE AS SEEN BY GAIA C. Reylé 1, A.C. Robin 1, M. Schultheis 1, S. Picaud 2 1 Observatoire de Besançon, CNRS UMR 6091, BP 1615, 25010 Besançon cedex, France 2 IAG/USP Departamento de

More information

Intense Star Formation in Nearby Merger Galaxies

Intense Star Formation in Nearby Merger Galaxies Intense Star Formation in Nearby Merger Galaxies 13 February 2009 Authors: Kelsey Johnson (University of Virginia), Sara Beck (Tel Aviv University), Aaron Evans (University of Virginia), Miller Goss (NRAO),

More information

Observing the Formation of Dense Stellar Nuclei at Low and High Redshift (?) Roderik Overzier Max-Planck-Institute for Astrophysics

Observing the Formation of Dense Stellar Nuclei at Low and High Redshift (?) Roderik Overzier Max-Planck-Institute for Astrophysics Observing the Formation of Dense Stellar Nuclei at Low and High Redshift (?) Roderik Overzier Max-Planck-Institute for Astrophysics with: Tim Heckman (JHU) GALEX Science Team (PI: Chris Martin), Lee Armus,

More information

dell Osservatorio 3, I Padova, Italy I Padova, Italy 98195, USA Wien, Austria 1. Introduction

dell Osservatorio 3, I Padova, Italy I Padova, Italy 98195, USA Wien, Austria 1. Introduction **Volume Title** ASP Conference Series, Vol. **Volume Number** **Author** c **Copyright Year** Astronomical Society of the Pacific Constraining Mass-Loss & Lifetimes of Low Mass, Low Metallicity AGB Stars

More information

Massimo Meneghetti 1, Elena Torri 1, Matthias Bartelmann 2, Lauro Moscardini 3, Elena Rasia 1 and Giuseppe Tormen 1,

Massimo Meneghetti 1, Elena Torri 1, Matthias Bartelmann 2, Lauro Moscardini 3, Elena Rasia 1 and Giuseppe Tormen 1, Mem. S.A.It. Vol. 73, 23 c SAIt 2002 Memorie della! "$# %&'()*+,(+ -. Massimo Meneghetti 1, Elena Torri 1, Matthias Bartelmann 2, Lauro Moscardini 3, Elena Rasia 1 and Giuseppe Tormen 1, 1 Dipartimento

More information

ASTR 1120 General Astronomy: Stars & Galaxies

ASTR 1120 General Astronomy: Stars & Galaxies ASTR 1120 General Astronomy: Stars & Galaxies!NNOUNCEMENTS HOMEWORK #6 DUE TODAY, by 5pm HOMEWORK #7 DUE Nov. 10, by 5pm Dark matter halo for galaxies Dark matter extends beyond visible part of the galaxy

More information

Review of Lecture 15 3/17/10. Lecture 15: Dark Matter and the Cosmic Web (plus Gamma Ray Bursts) Prof. Tom Megeath

Review of Lecture 15 3/17/10. Lecture 15: Dark Matter and the Cosmic Web (plus Gamma Ray Bursts) Prof. Tom Megeath Lecture 15: Dark Matter and the Cosmic Web (plus Gamma Ray Bursts) Prof. Tom Megeath A2020 Disk Component: stars of all ages, many gas clouds Review of Lecture 15 Spheroidal Component: bulge & halo, old

More information

Searching for Progenitors of Core-Collapse Supernovae

Searching for Progenitors of Core-Collapse Supernovae 1604 2004: SUPERNOVAE AS COSMOLOGICAL LIGHTHOUSES ASP Conference Series, Vol. 342, 2005 M. Turatto, S. Benetti, L. Zampieri, and W. Shea Searching for Progenitors of Core-Collapse Supernovae Schuyler D.

More information

Galaxies. CESAR s Booklet

Galaxies. CESAR s Booklet What is a galaxy? Figure 1: A typical galaxy: our Milky Way (artist s impression). (Credit: NASA) A galaxy is a huge collection of stars and interstellar matter isolated in space and bound together by

More information

Lecture 8: Stellar evolution II: Massive stars

Lecture 8: Stellar evolution II: Massive stars Lecture 8: Stellar evolution II: Massive stars Senior Astrophysics 2018-03-27 Senior Astrophysics Lecture 8: Stellar evolution II: Massive stars 2018-03-27 1 / 29 Outline 1 Stellar models 2 Convection

More information

CHAPTER 28 STARS AND GALAXIES

CHAPTER 28 STARS AND GALAXIES CHAPTER 28 STARS AND GALAXIES 28.1 A CLOSER LOOK AT LIGHT Light is a form of electromagnetic radiation, which is energy that travels in waves. Waves of energy travel at 300,000 km/sec (speed of light Ex:

More information

ASTR 1120 General Astronomy: Stars & Galaxies

ASTR 1120 General Astronomy: Stars & Galaxies ASTR 1120 General Astronomy: Stars & Galaxies!NNOUNCEMENTS HOMEWORK #6 DUE TODAY, by 5pm HOMEWORK #7 DUE Nov. 10, by 5pm Dark matter halo for galaxies REVIEW Dark matter extends beyond visible part of

More information

Searching for Other Worlds

Searching for Other Worlds Searching for Other Worlds Lecture 32 1 In-Class Question What is the Greenhouse effect? a) Optical light from the Sun is reflected into space while infrared light passes through the atmosphere and heats

More information

Lecture 30. The Galactic Center

Lecture 30. The Galactic Center Lecture 30 History of the Galaxy Populations and Enrichment Galactic Evolution Spiral Arms Galactic Types Apr 5, 2006 Astro 100 Lecture 30 1 The Galactic Center The nature of the center of the Galaxy is

More information

2) On a Hertzsprung-Russell diagram, where would you find red giant stars? A) upper right B) lower right C) upper left D) lower left

2) On a Hertzsprung-Russell diagram, where would you find red giant stars? A) upper right B) lower right C) upper left D) lower left Multiple choice test questions 2, Winter Semester 2015. Based on parts covered after mid term. Essentially on Ch. 12-2.3,13.1-3,14,16.1-2,17,18.1-2,4,19.5. You may use a calculator and the useful formulae

More information

Supernovae and the Accelerating Universe

Supernovae and the Accelerating Universe Supernovae and the Accelerating Universe Nicholas B. Suntzeff Mitchell Institute for Fundamental Physics Department of Physics & Astronomy Texas A&M University University of Texas/Austin Second Texas Cosmology

More information

Measurement of the stellar irradiance

Measurement of the stellar irradiance Measurement of the stellar irradiance Definitions Specific Intensity : (monochromatic) per unit area normal to the direction of radiation per unit solid angle per unit wavelength unit (or frequency) per

More information

The Milky Way Galaxy and Interstellar Medium

The Milky Way Galaxy and Interstellar Medium The Milky Way Galaxy and Interstellar Medium Shape of the Milky Way Uniform distribution of stars in a band across the sky lead Thomas Wright, Immanuel Kant, and William Herschel in the 18th century to

More information

PoS(IX EVN Symposium)019

PoS(IX EVN Symposium)019 abc, K.-H. Mack, a C.R. Benn, d R. Carballo, e D. Dallacasa, f a J.I. González-Serrano, g J. Holt h and F. Jiménez-Luján gi a INAF - Istituto di Radioastronomia, Via P. Gobetti 101, I-40129 Bologna, Italy

More information

Part 3: The Dark Energy

Part 3: The Dark Energy Part 3: The Dark Energy What is the fate of the Universe? What is the fate of the Universe? Copyright 2004 Pearson Education, published as Addison Weasley. 1 Fate of the Universe can be determined from

More information

Evidence for Supernova Light in all Gamma-Ray Burst Afterglow

Evidence for Supernova Light in all Gamma-Ray Burst Afterglow Clemson University TigerPrints Publications Physics and Astronomy Winter 12-13-2004 Evidence for Supernova Light in all Gamma-Ray Burst Afterglow A. Zeh Thüringer Landessternwarte Tautenburg S. Klose Thüringer

More information